infrared pinem developed by diffraction in 4d uemuem is photon-induced near-field electron...

6
Infrared PINEM developed by diffraction in 4D UEM Haihua Liu a , John Spencer Baskin a , and Ahmed H. Zewail a,1 a Physical Biology Center for Ultrafast Science and Technology, Arthur Amos Noyes Laboratory of Chemical Physics, California Institute of Technology, Pasadena, CA 91125 Contributed by Ahmed H. Zewail, January 8, 2016 (sent for review December 23, 2015; reviewed by Fabrizio Carbone, Archie Howie, and X. Sunney Xie) The development of four-dimensional ultrafast electron micros- copy (4D UEM) has enabled not only observations of the ultrafast dynamics of photonmatter interactions at the atomic scale with ultrafast resolution in image, diffraction, and energy space, but photonelectron interactions in the field of nanoplasmonics and nanophotonics also have been captured by the related technique of photon-induced near-field electron microscopy (PINEM) in im- age and energy space. Here we report a further extension in the ongoing development of PINEM using a focused, nanometer-scale, electron beam in diffraction space for measurements of infrared- light-induced PINEM. The energy resolution in diffraction mode is unprecedented, reaching 0.63 eV under the 200-keV electron beam illumination, and separated peaks of the PINEM electron-energy spectrum induced by infrared light of wavelength 1,038 nm (pho- ton energy 1.2 eV) have been well resolved for the first time, to our knowledge. In a comparison with excitation by green (519-nm) pulses, similar first-order PINEM peak amplitudes were obtained for optical fluence differing by a factor of more than 60 at the interface of copper metal and vacuum. Under high fluence, the nonlinear regime of IR PINEM was observed, and its spatial de- pendence was studied. In combination with PINEM temporal gat- ing and low-fluence infrared excitation, the PINEM diffraction method paves the way for studies of structural dynamics in re- ciprocal space and energy space with high temporal resolution. electron microscopy | diffraction | nanostructures | materials science | ultrafast dynamics S ince its invention in the 1930s by Knoll and Ruska (1), the electron microscope has become a powerful tool in the fields of physics, chemistry, materials, and biology. A great variety of techniques related to the electron microscope has been de- veloped in image, diffraction, and energy space (2, 3), with the spatial and energy resolutions of the transmission electron mi- croscope now reaching 0.5 Å with Cs corrector (4) and sub-100 meV with electron monochromators (5, 6), respectively. To these capabilities of spatial and energy resolution has been added the high resolution in the fourth dimension (time) by the development of four-dimensional ultrafast electron microscopy (4D UEM) (79), currently enabling nanoscale dynamic studies with temporal resolution that is 10 orders of magnitude better than the millisecond range of video-camera-rate recording in conventional microscopes. In 4D UEM, ultrafast time resolution is reached by using two separate but synchronized ultrashort laser pulses, one to generate a probing electron pulse by photoemission at the microscope cathode and the other to excite the specimen into a nonequilibrium state. The state of the specimen within the window of time of the probe pulse can be observed by recording the probe electron packet scattered from the specimen in any of the different working modes of the microscope, such as image and diffraction (10), energy spectrum (11), convergent beam (12), or scanning transmission electron microscopy (TEM) (13). Scanning the time delay between arrival of the pump and probe pulses at the specimen, which is controlled by a precise optical delay line, allows the evolution of the specimen to be traced. One of the important techniques developed in, and unique to, UEM is photon-induced near-field electron microscopy (PINEM) (14). PINEM has extended the capability of UEM to observation of lightelectron interactions near nanostructures or at an in- terface, which offers exciting prospects for the study of dynamics of photonics and plasmonics at the nanometer scale (15). The three-body interaction of photon, electron, and nanostructure relaxes momentum conservation and leads to efficient coupling between photons and electrons (16). In PINEM, an ultrashort optical pulse is used to excite evanescent electromagnetic fields near a nanostructure or at an interface. When the probe electron packet is in spatiotemporal overlap with these evanescent or scatter fields, some of its electrons can absorb/emit one or more scattered photons and then be detected by their contributions to displaced energy peaks in the electron energy spectrum. These displaced peaks appear as discrete sidebands to the zero-loss peak at separations given by the photon energy (hν) of the pump optical pulse. When using energy filtering to select for imaging only those electrons gaining energy, the resulting PINEM image reflects the strength and topology of the excited near field around the nanostructure or interface. The PINEM technique has been used to detect the evanescent near field surrounding a variety of structures with different ma- terials properties and different geometries, such as carbon nano- tubes (14), silver nanowires (14, 17), nanoparticles (16, 18), cells and protein vesicles (19), and several-atoms-thick graphene-layered steps (20). In addition, focused-beam PINEM has been used in scanning TEM mode to obtain induced near-field distributions for a copper grid bar (21), a nanometer gold tip (22), and a silver nanoparticle at the subparticle level (21). In a recent publication, three pulses, two optical and one electron, were introduced into the arsenal of techniques to gate the electron pulse and make its width only limited by the optical-pulse durations (23). Numerous general theoretical treatments (2428) have successfully described the phenomenon, with detailed treatments quantitatively repro- ducing many unique features of these multifaceted experimental observations (17, 2022). Despite the growing body of PINEM studies, almost all pre- viously published PINEM results were obtained in the image mode of the electron microscope using optical pulses with wavelengths of 500800 nm. An exception is a single unresolved Significance One of the most significant advances made is the ability to vi- sualize nanoscale objects and to determine their shapes (through imaging) and their atomic-scale structure (through diffraction). Furthermore, for optical excitation of the nanostructure the photon-induced near-field EM (PINEM) imaging results in the mapping of nanostructure plasmonics with high spatialtemporal resolution. Here, we show that PINEM can be exploited in the infrared region with greater sensitivity, and that diffraction can be implemented for structural dynamics. Just as importantly, the time resolution of ultrafast EM (UEM), which already far exceeds, by 10 orders of magnitude, that of conventional EM, may be further enhanced by IR PINEM and photon gating with the unprecedented energy resolution (0.63 eV) reported here for UEM. Author contributions: H.L., J.S.B., and A.H.Z. designed research; H.L. performed experi- ments; H.L. and J.S.B. analyzed data; and H.L., J.S.B., and A.H.Z. wrote the paper. Reviewers: F.C., École Polytechnique Fédérale de Lausanne; A.H., University of Cambridge; and X.S.X., Harvard University. The authors declare no conflict of interest. 1 To whom correspondence should be addressed. Email: [email protected]. www.pnas.org/cgi/doi/10.1073/pnas.1600317113 PNAS | February 23, 2016 | vol. 113 | no. 8 | 20412046 PHYSICS Downloaded by guest on August 1, 2020

Upload: others

Post on 06-Jul-2020

6 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Infrared PINEM developed by diffraction in 4D UEMUEM is photon-induced near-field electron microscopy (PINEM) (14). PINEM has extended the capability of UEM to observation of light–electron

Infrared PINEM developed by diffraction in 4D UEMHaihua Liua, John Spencer Baskina, and Ahmed H. Zewaila,1

aPhysical Biology Center for Ultrafast Science and Technology, Arthur Amos Noyes Laboratory of Chemical Physics, California Institute of Technology,Pasadena, CA 91125

Contributed by Ahmed H. Zewail, January 8, 2016 (sent for review December 23, 2015; reviewed by Fabrizio Carbone, Archie Howie, and X. Sunney Xie)

The development of four-dimensional ultrafast electron micros-copy (4D UEM) has enabled not only observations of the ultrafastdynamics of photon–matter interactions at the atomic scale withultrafast resolution in image, diffraction, and energy space, butphoton–electron interactions in the field of nanoplasmonics andnanophotonics also have been captured by the related techniqueof photon-induced near-field electron microscopy (PINEM) in im-age and energy space. Here we report a further extension in theongoing development of PINEM using a focused, nanometer-scale,electron beam in diffraction space for measurements of infrared-light-induced PINEM. The energy resolution in diffraction mode isunprecedented, reaching 0.63 eV under the 200-keV electron beamillumination, and separated peaks of the PINEM electron-energyspectrum induced by infrared light of wavelength 1,038 nm (pho-ton energy 1.2 eV) have been well resolved for the first time, toour knowledge. In a comparison with excitation by green (519-nm)pulses, similar first-order PINEM peak amplitudes were obtainedfor optical fluence differing by a factor of more than 60 at theinterface of copper metal and vacuum. Under high fluence, thenonlinear regime of IR PINEM was observed, and its spatial de-pendence was studied. In combination with PINEM temporal gat-ing and low-fluence infrared excitation, the PINEM diffractionmethod paves the way for studies of structural dynamics in re-ciprocal space and energy space with high temporal resolution.

electron microscopy | diffraction | nanostructures | materials science |ultrafast dynamics

Since its invention in the 1930s by Knoll and Ruska (1), theelectron microscope has become a powerful tool in the fields

of physics, chemistry, materials, and biology. A great variety oftechniques related to the electron microscope has been de-veloped in image, diffraction, and energy space (2, 3), with thespatial and energy resolutions of the transmission electron mi-croscope now reaching 0.5 Å with Cs corrector (4) and sub-100 meVwith electron monochromators (5, 6), respectively.To these capabilities of spatial and energy resolution has been

added the high resolution in the fourth dimension (time) by thedevelopment of four-dimensional ultrafast electron microscopy(4D UEM) (7–9), currently enabling nanoscale dynamic studieswith temporal resolution that is 10 orders of magnitude betterthan the millisecond range of video-camera-rate recording inconventional microscopes. In 4D UEM, ultrafast time resolutionis reached by using two separate but synchronized ultrashort laserpulses, one to generate a probing electron pulse by photoemissionat the microscope cathode and the other to excite the specimeninto a nonequilibrium state. The state of the specimen within thewindow of time of the probe pulse can be observed by recordingthe probe electron packet scattered from the specimen in any ofthe different working modes of the microscope, such as image anddiffraction (10), energy spectrum (11), convergent beam (12), orscanning transmission electron microscopy (TEM) (13). Scanningthe time delay between arrival of the pump and probe pulses atthe specimen, which is controlled by a precise optical delay line,allows the evolution of the specimen to be traced.One of the important techniques developed in, and unique to,

UEM is photon-induced near-field electron microscopy (PINEM)(14). PINEM has extended the capability of UEM to observationof light–electron interactions near nanostructures or at an in-terface, which offers exciting prospects for the study of dynamics

of photonics and plasmonics at the nanometer scale (15). Thethree-body interaction of photon, electron, and nanostructurerelaxes momentum conservation and leads to efficient couplingbetween photons and electrons (16). In PINEM, an ultrashortoptical pulse is used to excite evanescent electromagnetic fieldsnear a nanostructure or at an interface. When the probe electronpacket is in spatiotemporal overlap with these evanescent orscatter fields, some of its electrons can absorb/emit one or morescattered photons and then be detected by their contributions todisplaced energy peaks in the electron energy spectrum. Thesedisplaced peaks appear as discrete sidebands to the zero-losspeak at separations given by the photon energy (hν) of the pumpoptical pulse. When using energy filtering to select for imagingonly those electrons gaining energy, the resulting PINEM imagereflects the strength and topology of the excited near fieldaround the nanostructure or interface.The PINEM technique has been used to detect the evanescent

near field surrounding a variety of structures with different ma-terials properties and different geometries, such as carbon nano-tubes (14), silver nanowires (14, 17), nanoparticles (16, 18), cellsand protein vesicles (19), and several-atoms-thick graphene-layeredsteps (20). In addition, focused-beam PINEM has been used inscanning TEM mode to obtain induced near-field distributions fora copper grid bar (21), a nanometer gold tip (22), and a silvernanoparticle at the subparticle level (21). In a recent publication,three pulses, two optical and one electron, were introduced intothe arsenal of techniques to gate the electron pulse and make itswidth only limited by the optical-pulse durations (23). Numerousgeneral theoretical treatments (24–28) have successfully describedthe phenomenon, with detailed treatments quantitatively repro-ducing many unique features of these multifaceted experimentalobservations (17, 20–22).Despite the growing body of PINEM studies, almost all pre-

viously published PINEM results were obtained in the imagemode of the electron microscope using optical pulses withwavelengths of 500–800 nm. An exception is a single unresolved

Significance

One of the most significant advances made is the ability to vi-sualize nanoscale objects and to determine their shapes (throughimaging) and their atomic-scale structure (through diffraction).Furthermore, for optical excitation of the nanostructure thephoton-induced near-field EM (PINEM) imaging results inthe mapping of nanostructure plasmonics with high spatial–temporal resolution. Here, we show that PINEM can beexploited in the infrared region with greater sensitivity, andthat diffraction can be implemented for structural dynamics.Just as importantly, the time resolution of ultrafast EM (UEM),which already far exceeds, by 10 orders of magnitude, that ofconventional EM, may be further enhanced by IR PINEM andphoton gating with the unprecedented energy resolution (0.63eV) reported here for UEM.

Author contributions: H.L., J.S.B., and A.H.Z. designed research; H.L. performed experi-ments; H.L. and J.S.B. analyzed data; and H.L., J.S.B., and A.H.Z. wrote the paper.

Reviewers: F.C., École Polytechnique Fédérale de Lausanne; A.H., University of Cambridge;and X.S.X., Harvard University.

The authors declare no conflict of interest.1To whom correspondence should be addressed. Email: [email protected].

www.pnas.org/cgi/doi/10.1073/pnas.1600317113 PNAS | February 23, 2016 | vol. 113 | no. 8 | 2041–2046

PHYS

ICS

Dow

nloa

ded

by g

uest

on

Aug

ust 1

, 202

0

Page 2: Infrared PINEM developed by diffraction in 4D UEMUEM is photon-induced near-field electron microscopy (PINEM) (14). PINEM has extended the capability of UEM to observation of light–electron

PINEM spectrum for 1,038-nm excitation published from thislaboratory (25). Because the PINEM response of a material isgoverned by its optical properties and dimensions relative to thewavelength of light, excitation wavelength is an important pa-rameter largely remaining to be explored experimentally.Here we report the development of IR PINEM using excitation

at the wavelength of 1,038 nm (photon energy 1.2 eV). The spa-tial- and fluence-dependent behavior of well-resolved IR PINEMinduced at the edge of a copper grid bar is examined by combiningnanometer-scale convergent-beam electron diffraction and dif-fraction-mode detection for electron-energy spectroscopy with anunprecedented energy resolution down to 0.63 eV at 200 keV.Different e-beam size effects were compared for PINEM gener-ated by green and IR pump pulses. The spatial dependence of IRPINEM at the interface was studied at low-pulse fluence (linearregime) and high-pulse fluence (nonlinear regime). Diffraction ofa gold crystal film was observed using the energy-resolved PINEMelectrons produced by interaction with the scatter field of theadjacent copper grid edge. Notably, substantial PINEM peakamplitudes were achievable at dramatically lower fluence for IRpulses than for green pulses, opening up a possible path for studiesof photosensitive materials. This general accessibility of strongPINEM signals is of particular importance for our primary interestof ultrafast dynamics, for which PINEM photon gating has thepotential to vastly improve temporal resolution.All PINEM experiments reported here were performed on the

California Institute of Technology UEM-2 apparatus. The op-eration voltage on UEM-2 is 200 keV. The laser system usedemits a train of ∼220-fs pulses with wavelength of 1,038 nm, setto operate at a repetition rate of 1 MHz. The laser output wasfrequency-doubled two successive times to provide the 259-nmpulses used to generate the electron packet (probe beam) at the200-keV microscope photocathode source. The residual 1,038-nmand 519-nm optical pulses were each available for use as thePINEM pump beam to excite the near-field plasmons atthe interface. All of the experiments were carried out with po-larization set to be perpendicular to the interface and in thesingle-electron regime (8) to eliminate space-charge effects. Indiffraction mode, a camera length of 920 mm and a spectrometerentrance aperture of 1 mm were used to obtain a small collectionangle for better energy resolution.

PINEM: Image Mode and Diffraction ModeFig. 1 illustrates schematically PINEM measurements in imagemode and diffraction mode of the microscope. For standardPINEM studies in image mode (Fig. 1A), the electron beam isunfocused and spread on the specimen, although focused beamshave also been used. On the microscope view screen, a typicalbright-field TEM image of one carbon nanotube is shown. Byusing an electron-energy-spectrometer entrance aperture whichencompasses the nanotube area, the PINEM electron-energyspectrum is integrated over the entire area covered by the ap-erture when the optical and electron pulses are coincident on thespecimen. In energy-filtered imaging using only those electronsthat gain energy from photons, a PINEM image of the near fieldsurrounding the nanotube is obtained. The dependence of thenear-field signal on delay between the optical and electron pulsescan be extracted from the recorded PINEM energy spectrum orimage. The temporal response, as shown at the bottom of thefigure, represents a convolution of the temporal profile of theinduced field with those of both the optical and electron pulses.In the diffraction mode measurement illustrated here (Fig.

1B), the electron beam is focused to nanometer scale on a goldfilm in close proximity to a copper grid bar. On the view screen, atypical convergent-beam electron diffraction (CBED) pattern isobserved; diffraction spots are enlarged (12). By using a smallerspectrometer entrance aperture to cover one single diffractionspot, the PINEM diffraction and energy spectrum can be obtainedfrom only that chosen diffraction spot. In this concept of PINEMdiffraction, the PINEM electrons exchange energy with the opticalfield scattered from the copper while being diffracted from ananometer scale area of the gold specimen. This makes it possibleto use a PINEM signal to study structural dynamics of a local areaof a specimen of arbitrary structure and optical properties. Thetemporal confinement of the PINEM electrons can thus beexploited for greatly improved time resolution in diffractionstudies (23, 29).Another advantage of recording PINEM in diffraction mode is

an improvement in the energy resolution in UEM operated at200 keV, which reaches the 0.63-eV value. This improvement,attributed to the reduced dispersion effect of electron scatteringin different directions, makes it easy to observe well-resolvedPINEM peaks separated by the IR photon energy of 1.2 eV, asshown by the IR PINEM electron-energy spectrum in Fig. 1B.The PINEM signal dependence on interpulse time delay can

Fig. 1. Variant implementations of PINEM in UEM.(A) PINEM in image mode using a parallel electronbeam and green optical pulses. The nanostructurespecimen is seen as a bright-field image on theview screen and as an energy-filtered PINEM imageor PINEM electron-energy spectrum after the spec-trometer. The time-delay dependence of the PINEMsignal shown at the bottom can be extracted from aseries of spectra or images. (B) PINEM in diffractionmode using a focused nanometer electron beam andIR optical pulses. The ultrafast electron pulses arefocused near the interface between a copper gridand gold film. A typical CBED pattern is observed onthe view screen, and a small spectrometer entranceaperture is used to select a single diffraction spot forenergy dispersion. After the spectrometer, an en-ergy-filtered PINEM diffraction image or electron-energy spectrum is recorded. For a description of thetime plot at the bottom, see A.

2042 | www.pnas.org/cgi/doi/10.1073/pnas.1600317113 Liu et al.

Dow

nloa

ded

by g

uest

on

Aug

ust 1

, 202

0

Page 3: Infrared PINEM developed by diffraction in 4D UEMUEM is photon-induced near-field electron microscopy (PINEM) (14). PINEM has extended the capability of UEM to observation of light–electron

again be extracted from the electron-energy spectrum (as de-scribed above for image mode) and is shown in Fig. 1B, Bottom.

PINEM: e-Beam Size and b-Material EffectThe distinction between the two types of experiment of Fig. 1 isfurther illustrated in Fig. 2 A and B. Fig. 2A shows an unfocusede-beam, spread to a diameter of about 1.7 μm, centered on theinterface between vacuum (Upper Left) and a copper grid (LowerRight), indicated by the solid red line. The dashed circle repre-sents the total e-beam illumination area, with no electronspassing through the copper grid due to its thickness (20-μmtypical value). Under this e-beam condition, the resulting PINEMelectron-energy spectrum is the integration of electron–photoninteractions over the entire beam cross-section in the vacuumregion. For such a large beam width, the PINEM response is veryinhomogeneous, resulting in measurement of the average signalover the illuminated area.In contrast, in Fig. 2B a series of three e-beams focused to

nanometer size (33-nm FWHM) is shown at the same imagescale as Fig. 2A. In this case the resulting PINEM is contrib-uted by the electron–photon interaction localized over a smallrange of b (impact parameter or distance from the interface)and all electrons are subject to a spatially homogeneous scat-tered light field. When the nanometer e-beam is scanned awayfrom the grid along the green arrow (increasing b), the spatial profileof PINEM induced by the evanescent near field at the interface canbe obtained, as done previously for green excitation (21).Fig. 2 C–F compares PINEM spectra recorded in diffraction

mode at two pump wavelengths and two e-beam diameters(d), all with similar amplitudes for their n = ±1 PINEM peaks(ΔE = ± hν). The probe beams in each case pass close to thecopper edge (b ∼ d/2) in vacuum and the temporal overlap be-tween electron and photon pulses is maximized (i.e., at delay t = 0).Shown in Fig. 2C is the PINEM electron-energy spectrum excitedby green (519-nm) optical pulses at a fluence of 2.7 mJ/cm2. Theenergy loss is referenced to the original zero-loss energy of theincident electron beam. The e-beam size used is 130 nm, obtainedby controlling the condenser lens strength and using a smallercondenser aperture. Because there is no substrate there are no

diffraction peaks; only the central directed beam was selectedfor dispersion in the spectrometer by a small entrance aperture.The discrete peaks of the PINEM energy spectrum at a sepa-ration of 2.4 eV are very well resolved, and up to 6-photonabsorption and emission can be observed with the amplitude ofeach successive peak decreasing by a factor of ∼2. The firstgain/loss peak shows a 48% amplitude ratio relative to that ofthe zero-loss peak (ZLP).For comparison, at the same interface and using the same

e-beam size and position, Fig. 2D shows the PINEM electron-energy spectrum for 1,038-nm excitation at a fluence of 0.04 mJ/cm2,also with n = ±1 PINEM peak amplitudes 48% of the ZLP.However, in this case only two gain/loss peaks, separated by an energyof 1.2 eV, are seen and well resolved. Thus, for the same experi-mental conditions and the same relative amplitude of the n = 1 peaksobtained for green and IR, the IR fluence is a factor of greater than60 times lower. Similarly, in Fig. 2 E and F, green- and IR-inducedPINEM spectra are compared for a smaller e-beam size of 33 nmadjacent to the same copper–vacuum interface. The greenfluence (1.6 mJ/cm2) is again much higher (now a factor of ∼4)than that of the IR (0.37 mJ/cm2) to produce similar n = 1 peakamplitude ratios.Comparisons between Fig. 2 C and E and between Fig. 2 D

and F show the effect of change in e-beam size (accompaniedby a change in impact parameter in this case) for PINEM at thecopper interface excited by the same light pulses. Similar PINEMgenerated by lower excitation fluence for the smaller e-beamsize (33 nm) than for the larger size (130 nm) is consistentwith the weak interaction limit of PINEM theory and a nearfield decaying away from the interface. This is the trend seenfor green excitation, but not for IR excitation (see below).

PINEM in Diffraction Mode: Energy ResolutionFig. 3A is an expanded view of the IR PINEM spectrum ofFig. 2D, showing the two discrete gain and loss peaks. Thezero-loss energy spectrum obtained under the same condi-tions, but at t = −3 ps (before the optical pump pulse arrival)is shown in Fig. 3B. The energy resolution is 0.63 eV, the bestenergy resolution reported in UEM at 200 keV. The energy

Fig. 2. (A) Unfocused e-beam experiments. The unfocused beam illuminates the interface between vacuum and copper grid. The dashed circle shows thee-beam coverage area (1.7 μm in diameter). (Scale bar, 500 nm.) (B) Focused e-beam experiments. The e-beam is focused to nanometer size (33-nm FWHM) atthe interface between the vacuum and Cu grid. The e-beam is scanned away from the grid along the green arrow to obtain the distance dependence of thenear field at the interface. (Scale bar, 500 nm.) The dashed circle shows the unfocused beam size from A for comparison. (C–F) PINEM electron-energy spectraobtained at t = 0 in diffraction mode at the interface between vacuum and Cu grid using two different e-beam sizes for two different wavelengths of theoptical pulse. The energy loss is referenced to the original zero-loss energy of the incident electron beam. All of the electron-energy spectra have similar firstPINEM peak ratio relative to the ZLP. (C) e-beam size 130 nm, green light (λ = 519 nm) with excitation fluence of 2.7 mJ/cm2. (D) e-beam size 130 nm, infraredlight (λ = 1,038 nm) with excitation fluence of 0.04 mJ/cm2. (E) e-beam size 33 nm, green, 1.6 mJ/cm2. (F) e-beam size 33 nm, infrared, 0.37 mJ/cm2.

Liu et al. PNAS | February 23, 2016 | vol. 113 | no. 8 | 2043

PHYS

ICS

Dow

nloa

ded

by g

uest

on

Aug

ust 1

, 202

0

Page 4: Infrared PINEM developed by diffraction in 4D UEMUEM is photon-induced near-field electron microscopy (PINEM) (14). PINEM has extended the capability of UEM to observation of light–electron

resolution in the microscope is measured from the FWHM ofthe ZLP. The energy resolution is mainly determined by thewidth of the energy spread provided by the electron source,which is typically between 1 and 2 eV for the tungsten filamentor heated LaB6 source and 0.5–1 eV for a Schottky or coldfield-emission source without electron monochromator (30).In pulsed UEM systems, most of which use thermionic sour-ces, resolution is typically above 1 eV at an operating energyof 200 keV (31).Also affecting the energy resolution are aberrations of the

electron spectrometer, which can be almost eliminated by curv-ing the polepieces of the magnetic prism and by using weakmultipole lenses for fine tuning (30). However, in actual exper-iments, the electrons are scattered by the specimen in differentdirections, especially when there is strong diffraction in a crys-talline specimen. Due to inherent characteristics of the magneticprism, electrons scattered in different directions hit differenttargets on the detector in energy space after passing through thespectrometer, which decreases the energy resolution, especiallyin image mode. In fact, when recording electron-energy spectra,it is not necessary to collect all of the scattered electrons, part ofwhich can be filtered out without harming the results.In the spectrometer, there are three important parameters

for optimization of the experimental conditions: the entranceaperture angle (γ), the object size, and the maximum collectionangle (β). In diffraction mode, these three parameters can beadjusted unambiguously and independently; γ is controlled bythe spectrometer entrance aperture, object size is determinedby the selected area aperture, or by the size of the illuminatedarea, and β can be adjusted by the camera length depending onthe size of the spectrometer entrance aperture. In our focusede-beam diffraction case, the object size is limited by thenanometer-scale area illuminated by the focused electronbeam. To improve energy resolution, a choice of suitable sizeof the spectrometer entrance aperture and camera length en-sures that only a single diffraction spot, consisting of electronsscattered in the same direction, is selected to form the electron

energy spectrum. The maximum collection angle β used here isabout 7 mrad.

IR PINEM: Spatial and Temporal Dependence in theLinear RegimeTo observe the spatial localization of electron–photon inter-actions in the PINEM effect excited by IR pulses at the in-terface of vacuum and Cu grid, the focused e-beam with a sizeof 33 nm was scanned into vacuum away from the Cu grid alongthe green arrow in Fig. 2B. A low IR optical pulse fluence of0.04 mJ/cm2 was used to remain in the linear regime. As shownin Fig. 3C, the spatial dependence of the amplitudes of first-and second-order PINEM peaks decays quasi-exponentially with b.The values shown are direct amplitude measurements in spectrarecorded with a constant exposure and electron flux. The solidlines are exponential fits, corresponding to decay constants of∼164 nm and 60 nm for n = 1 and n = 2, respectively. Fig. 3Dshows the time dependence of the first- and second-order peaksextracted from time-resolved PINEM energy spectra obtained atthe interface using e-beam size of 130 nm at an IR pulse fluence of0.04 mJ/cm2. The Gaussian fits indicated by the solid line giveFWHM values of 911 fs for n = 1 and 642 fs for n = 2.For understanding of the expected topology of the PINEM

signal at the copper grid edge, it is pertinent to give here a briefaccount of previous theoretical descriptions relevant to the sit-uation. The scattering of the electron packet traveling throughthe scattered electromagnetic wave of a nanoparticle or an in-terface induced by an optical pulse can be treated in the regimeof the time-dependent Schrödinger equation (25). The theoret-ical study has shown that PINEM can be related to opticallydriven charge-density distributions of nanoparticle plasmons(28). The PINEM intensity is approximately proportional to theabsolute square of the field integral ~F0 (i.e., IPIN ∝

��~F0��2). For the

case of an infinite thin strip with axis perpendicular to boththe propagation direction and polarization direction of the in-cident light pulse, as represented in dark gray in Fig. 3C (Inset), a

Fig. 3. (A) PINEM electron-energy spectrum obtained at the interface between the vacuum and Cu grid using an e-beam size of 130 nm in diffraction mode,under IR (1,038 nm) excitation fluence of 0.04 mJ/cm2. (B) The zero-loss energy spectrum obtained in diffraction mode, at t = −3 ps showing the energyresolution of 0.63 eV determined by the Gaussian fit (solid line). (C) Dependence of PINEM peak amplitudes on distance away from the grid–vacuum interface. Thee-beam (33 nm) was scanned along the green arrow in Fig. 2B. The solid lines show exponential fits, yielding damping constants of 164 nm and 60 nm, respectively forthe first- and second-order PINEM peaks. (Inset) Illustration of the geometry and x coordinate of the thin strip (dark gray) to which Eq. 1 applies. The light-gray blockrepresents schematically the actual Cu grid bar used in our experiments. (D) Time-delay dependence of the first and second PINEM peak amplitudes. The solid lines areGaussian fits (FWHM: 910 fs, 640 fs). (E) Bright-field UEM image showing the interface of gold film and Cu grid under unfocused e-beam illumination. (Scale bar, 1 μm.)(F) [11 0] zone axis, CBED pattern from an 11-nm-thick gold single-crystal film using a focused e-beam (300 nm). (Scale bar, 5 nm−1.)

2044 | www.pnas.org/cgi/doi/10.1073/pnas.1600317113 Liu et al.

Dow

nloa

ded

by g

uest

on

Aug

ust 1

, 202

0

Page 5: Infrared PINEM developed by diffraction in 4D UEMUEM is photon-induced near-field electron microscopy (PINEM) (14). PINEM has extended the capability of UEM to observation of light–electron

near-field approximation of the field integral derived in theweak-interaction limit has been given as (20)

~F0 ≈−i~E0χb  h2

ne−Δke jx−ðw=2Þj − e−Δke jx+ðw=2Þj

o, [1]

where E0 is the amplitude of the incident electric field, h is the stripheight, w the strip width, Δke the momentum change of the elec-tron, x the distance normal to the axis of the strip measured fromthe center of the strip width, and χb ≈ ½ðw+ hÞ=ðw+ «hÞ�ð«− 1Þ,where χb is the material susceptibility and « the dielectric function.From Eq. 1, the PINEM intensity is dependent on material prop-erties, shape and size of the strip, the polarization and intensity ofthe optical pulse, and location. The decay length in the exponentialterms is (Δke)−1, where Δk« =ωp=ν« for light of angular frequencyωp and an electron with velocity υe. This length is simply propor-tional to the wavelength of incident light; for 200-keV electronsand 1,038-nm light, (Δke)−1 is 115 nm. Thus, for a strip much widerthan the wavelength of light, the second term in Eq. 1 disappears,and the amplitude of the PINEM field is seen to decrease expo-nentially with distance away from the strip edge (located at w/2).Although Eq. 1 is an approximation derived for a thin strip (on

the nanometer scale), it provides qualitative agreement with thePINEM signal behavior shown in Fig. 3C for the case of the coppergrid bar at low fluence. However, the height of the grid is in factmuch greater than the wavelength of light. This micrometer extentof the specimen, represented by the light-gray slab in Fig. 3C (Inset),may introduce phase differences and inhomogeneity in the fieldinteracting with the electrons, and such effects must be properlyconsidered for a quantitative treatment of the signals observed here.The above comments about the effect of the thickness of the

grid bar are relevant to the comparison of Fig. 2 D and F also. Thevery large difference in fluence for similar PINEM signals whenreducing the e-beam size is not expected from consideration of Eq.1, but the thick grid and its relation to the wavelength of light maybe responsible for the difference in IR and green behavior. Futureexperiments should examine the wavelength, fluence, and e-beamsize dependence independently and for different structures.Fig. 3E shows a bright-field image of the interface between a

gold film and Cu grid illuminated by an unfocused e-beam. Atypical CBED pattern of [1 1 0] zone axis gold single crystal is

given in Fig. 3F, which was obtained at the interface using a300-nm e-beam. The IR PINEM electron-energy spectrum canbe obtained from each single diffraction spot or the centraldirected spot by using a suitably sized spectrometer entrance ap-erture. The study of structural dynamics for an arbitrary substrate athigh temporal resolution is possible in this PINEM configurationusing the power of PINEM photon gating (23, 29). In this tech-nique, an optical PINEM pump pulse is fixed in time with respect tothe electron pulse, thus gating a pulse of PINEM electrons of du-ration controlled by the optical pulse length. This subset of elec-trons, detected separately by energy filtering, becomes the probefor dynamics initiated by a second time-delayed ultrafast opticalpulse focused on the specimen. The achievable time resolution isin this case limited only by the optical pulse lengths. Note alsothat there is no restriction on the type of specimen that may bestudied in this scheme because the structural and material re-quirements for the generation of the PINEM signal are im-posed exclusively on the grid bar.

IR PINEM: Nonlinear PhenomenaTo further characterize the IR PINEM of the copper grid, the IRpump intensity was increased beyond the linear range. Fig. 4 Aand B shows two IR PINEM electron-energy spectra recorded indiffraction mode using a 33-nm e-beam under excitation fluenceof 0.08 mJ/cm2 and 2.2 mJ/cm2, respectively. As shown in Fig. 4Ain the low-fluence regime, only weak first-order absorption/emission peaks are observed in the PINEM energy spectrum.However, in the high-fluence regime of Fig. 4B, the PINEM energyspectrum shows multiple discrete sideband peaks with up to 13 netphotons absorbed and emitted. In contrast to PINEM energyspectra in the linear regime, the peak amplitudes in the orders fromn = ±2 to ±4 are even higher than the zero-peak amplitude, whichis called an “inverted distribution” behavior (25).Fig. 4C depicts the nonlinear behavior of spatial dependence of

IR PINEM peaks up to the 10th order obtained by e-beam (33 nm)scanning away from the interface between vacuum and Cu grid.The excitation fluence used here is 2.2 mJ/cm2. As for Fig. 3C, thepeak amplitudes are measured directly from spectra recorded un-der fixed conditions. Unlike the exponential decay of PINEM peakamplitude of the first- and second order in the linear regime (Fig.3C), the nonmonotonic behavior of peak amplitude from order n =±1 to order n = ±6 is very clear, in which the distance from the grid

Fig. 4. Nonlinear IR PINEM at the interface between the vacuum and Cu grid. (A and B) IR PINEM electron-energy spectra recorded in diffractionmode using a 33-nme-beam under different IR excitation fluence of (A) 0.08 mJ/cm2 and (B) 2.2 mJ/cm2. (C) Distance dependence of IR PINEM peaks up to the 10th order obtained byscanning the 33-nm e-beam away from the interface under excitation fluence of 2.2 mJ/cm2. (D) The energy resolution is 0. 7 eV measured from the zero-loss energyspectrum. (E) Time-and energy dependence of PINEM electron-energy spectra obtained at the interface for 130-nm e-beam and IR fluence of 1.25 mJ/cm2.

Liu et al. PNAS | February 23, 2016 | vol. 113 | no. 8 | 2045

PHYS

ICS

Dow

nloa

ded

by g

uest

on

Aug

ust 1

, 202

0

Page 6: Infrared PINEM developed by diffraction in 4D UEMUEM is photon-induced near-field electron microscopy (PINEM) (14). PINEM has extended the capability of UEM to observation of light–electron

of maximum peak amplitude decreases with increasing order ofabsorption/emission. The energy resolution used in spatial de-pendence experiments is 0.7 eV, as shown in Fig. 4D.Previous theoretical study predicted that in the weak-

interaction limit the temporal dependence is precisely the con-volution of incident electron packet and photon pulse profile,whereas for increasing interaction strengths broadening of thetemporal width and deviation from Gaussian behavior is the directresult of nonlinear effects (32). The probability amplitudes for thePINEM peaks of each order n depend on the nth-order Besselfunction of the first kind Jn(Ω), where Ω= eF0=ðZωÞ (25). At highfield strength, the Bessel function has the asymptotic form

JnðΩÞ=ffiffiffiffiffiffiffi2πΩ

rcos

�Ω−

nπ2−π

4

�. [2]

Thus, as the field in the high-fluence regime varies in spaceor time depending on the nature of the PINEM field, theprobabilities display an oscillatory behavior. For the focusede-beam, due to well-defined impact parameter b, such oscilla-tions with field variation in space are not averaged out (25). Therise and fall of PINEM amplitude with b in Fig. 4C is a conse-quence of this cosine-like behavior with respect to the changingstrength of the scattered light field. Such oscillatory or nonmono-tonic behavior of PINEM, seen in Fig. 4C and elsewhere (21, 22),represents the nature of electron–photon coupling in PINEM, co-herent versus incoherent (25).Fig. 4E gives the time- and energy dependence of PINEM

electron-energy spectra obtained at the interface of copper gridusing an e-beam of 130 nm, IR fluence of 1.25 mJ/cm2. It showssymmetrical energy dependence for all orders and asymmetricaltime dependence at low order. Such low-order time asymmetrywas commented on by Kirchner et al. (33) in reference to time–energy patterns like those shown here, but using shorter opticalpulses and 25 keV for the electrons. It is attributed to weak

temporal asymmetry of the optical pulses that is highly magnifiedby the strong nonlinear PINEM response.

OutlookUnder focused electron beam conditions, PINEM experimentswere carried out at an interface between vacuum and Cu grid indiffraction mode. The energy resolution remarkably reached the0.63-eV value, which makes the observation of resolved PINEMpeaks excited by IR (1,038 nm) possible. By comparing thePINEM effects between green and IR, the near-field enhance-ment in PINEM at the interface is shown to be much moresensitive to IR pulses, and this provides one promising way toinvestigate near-field effects using more gentle optical pulseswithout damaging the specimen before useful information isobtained, especially for biological sample visualization. PINEMdiffraction was realized, to our knowledge, for the first time,which opens the door to studies of structural dynamics of ma-terials by the PINEM-gating technique to reach a higher tem-poral resolution. The nonlinear PINEM effects by IR at highfluence will be further studied by combining the theoreticalcalculation to determine the nature of the multiple photon–electron coupling interactions, coherent or incoherent.Going beyond the scope of this work, these results for IR ir-

radiation and the earlier visible excitation (14, 15, 25) provide asuggestion of the opportunities that may arise when opticalscanning through resonances is made possible using high-power,broadly tunable lasers. Further PINEM enhancement is expected,with a potential for additional applications in spectroscopy andplasmonics, already the subject of studies at high spatial and en-ergy resolutions (34–38). The direct excitation of very low-energymodes, such as phonons and bond vibrations, may enable selectedmode imaging in this domain of resonant excitation (36).

ACKNOWLEDGMENTS. This work was supported by the National ScienceFoundation Grant DMR-0964886 and the Air Force Office of Scientific ResearchGrant FA9550-11-1-0055 for research conducted in The Gordon and BettyMoore Center for Physical Biology at the California Institute of Technology.

1. Knoll M, Ruska E (1932) Das Elektronenmikroskop. Z Phys 78(5):318–339.2. Hawkes PW, Spence JCH, eds (2007) Science of Microscopy (Springer, New York).3. Williams DB, Carter CB (2009) Transmission Electron Microscopy - A Textbook for

Materials Science (Springer Science Business Media, New York).4. Kisielowski C, et al. (2008) Detection of single atoms and buried defects in three di-

mensions by aberration-corrected electron microscope with 0.5-A information limit.Microsc Microanal 14(5):469–477.

5. Tiemeijer PC (1999) Measurement of Coulomb interactions in an electron beammonochromator. Ultramicroscopy 78(1-4):53–62.

6. KimotoK, KothleitnerG, GroggerW,Matsui Y, Hofer F (2005)Advantages of amonochromatorfor bandgap measurements using electron energy-loss spectroscopy. Micron 36(2):185–189.

7. Zewail AH, Thomas JM (2009) 4D Electron Microscopy: Imaging in Space and Time(Imperial College Press, London).

8. Zewail AH (2010) Four-dimensional electron microscopy. Science 328(5975):187–193.9. Zewail AH (2014) 4D Visualization of Matter (Imperial College Press, London).10. Barwick B, Park HS, Kwon OH, Baskin JS, Zewail AH (2008) 4D imaging of transient

structures andmorphologies in ultrafast electronmicroscopy. Science 322(5905):1227–1231.11. Carbone F, Kwon OH, Zewail AH (2009) Dynamics of chemical bonding mapped by

energy-resolved 4D electron microscopy. Science 325(5937):181–184.12. Yurtsever A, Zewail AH (2009) 4D nanoscale diffraction observed by convergent-beam

ultrafast electron microscopy. Science 326(5953):708–712.13. Ortalan V, Zewail AH (2011) 4D scanning transmission ultrafast electron microscopy:

Single-particle imaging and spectroscopy. J Am Chem Soc 133(28):10732–10735.14. Barwick B, Flannigan DJ, Zewail AH (2009) Photon-induced near-field electron mi-

croscopy. Nature 462(7275):902–906.15. Barwick B, Zewail AH (2015) Photonics and plasmonics in 4D ultrafast electron mi-

croscopy. ACS Photonics 2(10):1391–1402.16. Yurtsever A, Zewail AH (2012) Direct visualization of near-fields in nanoplasmonics

and nanophotonics. Nano Lett 12(6):3334–3338.17. Piazza L, et al. (2015) Simultaneous observation of the quantization and the in-

terference pattern of a plasmonic near-field. Nat Commun 6:6407.18. Yurtsever A, Baskin JS, Zewail AH (2012) Entangled nanoparticles: Discovery by vi-

sualization in 4D electron microscopy. Nano Lett 12(9):5027–5032.19. Flannigan DJ, Barwick B, Zewail AH (2010) Biological imaging with 4D ultrafast

electron microscopy. Proc Natl Acad Sci USA 107(22):9933–9937.20. Park ST, Yurtsever A, Baskin JS, Zewail AH (2013) Graphene-layered steps and their

fields visualized by 4D electron microscopy. Proc Natl Acad Sci USA 110(23):9277–9282.

21. Yurtsever A, van der Veen RM, Zewail AH (2012) Subparticle ultrafast spectrum im-aging in 4D electron microscopy. Science 335(6064):59–64.

22. Feist A, et al. (2015) Quantum coherent optical phase modulation in an ultrafasttransmission electron microscope. Nature 521(7551):200–203.

23. Hassan MT, Liu H, Baskin JS, Zewail AH (2015) Photon gating in four-dimensionalultrafast electron microscopy. Proc Natl Acad Sci USA 112(42):12944–12949.

24. García de Abajo FJ, Asenjo-Garcia A, KociakM (2010) Multiphoton absorption and emissionby interaction of swift electrons with evanescent light fields. Nano Lett 10(5):1859–1863.

25. Park ST, Lin M, Zewail AH (2010) Photon-induced near-field electron microscopy(PINEM): Theoretical and experimental. New J Phys 12:123028.

26. Howie A (2011) Photon interactions for electron microscopy applications. Eur Phys JAppl Phys 54(3):33502.

27. Asenjo-Garcia A, Garcia de Abajo FJ (2013) Plasmon electron energy-gain spectros-copy. New J Phys 15:103021.

28. Park ST, Zewail AH (2014) Photon-induced near-field electron microscopy: Mathe-matical formulation of the relation between the experimental observables and theoptically driven charge density of nanoparticles. Phys Rev A 89(1):013851.

29. Park ST, Zewail AH (2012) Enhancing image contrast and slicing electron pulses in 4Dnear field electron microscopy. Chem Phys Lett 521:1–6.

30. Egerton RF (2009) Electron energy-loss spectroscopy in the TEM. Rep Prog Phys 72:016502.31. Piazza L, et al. (2013) Design and implementation of a fs-resolved transmission

electron microscope based on thermionic gun technology. Chem Phys 423:79–84.32. Plemmons DA, Tae Park S, Zewail AH, Flannigan DJ (2014) Characterization of fast

photoelectron packets in weak and strong laser fields in ultrafast electron micros-copy. Ultramicroscopy 146:97–102.

33. Kirchner FO, Gliserin A, Krausz F, Baum P (2014) Laser streaking of free electrons at25 keV. Nat Photonics 8:52–57.

34. Matz R, Lueth H (1981) Conduction-band surface plasmons in the electron-energy-lossspectrum of GaAs(110). Phys Rev Lett 46(7):500–503.

35. Garcia de Abajo FJ (2014) Graphene plasmonics: Challenges and opportunities. ACSPhotonics 1(3):135–152.

36. Howie A (2015) Stimulated excitation electron microscopy and spectroscopy.Ultramicroscopy 151:116–121.

37. Krivanek OL, et al. (2014) Vibrational spectroscopy in the electron microscope. Nature514(7521):209–212.

38. Schaffer B, Hohenester U, Truegler A, Hofer F (2009) High-resolution surface plasmonimaging of gold nanoparticles by energy-filtered transmission electron microscopy.Phys Rev B 79(4):041401.

2046 | www.pnas.org/cgi/doi/10.1073/pnas.1600317113 Liu et al.

Dow

nloa

ded

by g

uest

on

Aug

ust 1

, 202

0