Brought to you by:
Review Article

Experimental methods of molecular matter-wave optics

, and

Published 2 August 2013 © 2013 IOP Publishing Ltd
, , Citation Thomas Juffmann et al 2013 Rep. Prog. Phys. 76 086402 DOI 10.1088/0034-4885/76/8/086402

0034-4885/76/8/086402

Abstract

We describe the state of the art in preparing, manipulating and detecting coherent molecular matter. We focus on experimental methods for handling the quantum motion of compound systems from diatomic molecules to clusters or biomolecules.

Molecular quantum optics offers many challenges and innovative prospects: already the combination of two atoms into one molecule takes several well-established methods from atomic physics, such as for instance laser cooling, to their limits. The enormous internal complexity that arises when hundreds or thousands of atoms are bound in a single organic molecule, cluster or nanocrystal provides a richness that can only be tackled by combining methods from atomic physics, chemistry, cluster physics, nanotechnology and the life sciences.

We review various molecular beam sources and their suitability for matter-wave experiments. We discuss numerous molecular detection schemes and give an overview over diffraction and interference experiments that have already been performed with molecules or clusters.

Applications of de Broglie studies with composite systems range from fundamental tests of physics up to quantum-enhanced metrology in physical chemistry, biophysics and the surface sciences.

Nanoparticle quantum optics is a growing field, which will intrigue researchers still for many years to come. This review can, therefore, only be a snapshot of a very dynamical process.

Export citation and abstract BibTeX RIS

Table of symbols

bgrating thickness
C4Casimir–Polder interaction constant
dgrating period
Δdvariation of grating period
$\vert e\vec{p_{0}}+ \hslash \vec{k}\rangle$ two-level atom in its excited state, after photon absorption
Eiionization energy
Eelectric field
Fforce
G1, G2, G3gratings 1, 2 and 3
|g, p0two-level atom in its ground state moving with momentum p0
hPlanck's constant
$\hslash$ h/2π
Jn(φ)Bessel function of nth order with modulation index φ
kphoton wavenumber, 2π/λ
kBBoltzmann's constant
lclongitudinal (spectral) coherent length
Llength
LTTalbot length
mmass
nindex, n ∈ N
neelectron number density
Nnumber of grating openings
polinear forward-directed momentum of a particle
pininitial momentum of a particle
Δpmomentum change
PLlaser power
rinteraction distance molecule/grating wall
Ttemperature
Ttranstranslational temperature
Trotrotational temperature
Tvibvibrational temperature
vmodulus of the velocity $v=\vert \vec{v}{\vert}$
v0most probable velocity
vlonglongitudinal velocity
Δvspread of longitudinal velocity
vtranstransverse velocity
VCPCasimir–Polder interaction potential
wxwaist of laser beam in the x-direction
wywaist of laser beam in the y-direction
xposition
xcCasimir cut-off distance
Δxsource width
Δx3increment of position of G3
Xctransverse (spatial) coherence width
αstatic (frequency-independent) polarizability
αoptoptical (frequency-dependent, dynamic) polarizability
βndiffraction angle of nth diffraction order
ηionionization efficiency
ηsesecondary electron yield
Φphase of the matter wave
ΔΦphase shift of the matter wave
λdBde Broglie wavelength
ΔλdBspread of de Broglie wavelength distribution
λLwavelength of light
ψwave function
σion(E)molecular ionization cross section
τtransit time of particle through laser waist

Table of acronyms, alphabetically sorted

2PItwo-photon ionization
AFMatomic force microscope
BECBose–Einstein condensate
EMCCDelectron multiplying charge coupled device
ESIelectrospray ionization
FWHMfull-width at half-maximum
IOPInstitute of Physics, UK
KDTLIKapitza–Dirac–Talbot–Lau interferometer/interferometry
LDlaser desorption
LIADlaser-induced acoustic desorption
LIFTlaser-induced forward transfer
MALDImatrix-assisted laser desorption ionization
MCPmulti-channel plate
MUPImulti-photon ionization
MZIMach–Zehnder interferometer
NSOMnear-field scanning optical microscope
OTIMAoptical time-domain ionizing matter-wave interferometer
PSFpoint spread function
QMSquadrupole mass spectrometer/spectroscopy
R2PIresonant two-photon ionization
REMPIresonantly enhanced multi-photon ionization
SPIsingle-photon ionization
SQUIDsuperconducting quantum interference device
SSPDoriginally: superconducting single-photon detectors; now and here also: superconducting single-particle detectors
STMscanning tunnelling microscope
TLITalbot–Lau interferometer/interferometry
TOF-MS or TOFtime-of-flight-mass spectroscopy
UV/VUVultraviolet (10–400 nm)/vacuum ultraviolet (10–200 nm)

1. Introduction: Why matter-wave optics with clusters and molecules?

The physics of matter waves has intrigued a large scientific community over the past few decades [14]. It is both at the core of many non-intuitive but fundamental quantum phenomena and it has become a central element of modern devices, which use tiny shifts of the quantum waves to measure material structures [6, 7], properties [810] or external forces [11]. In this report, we focus specifically on the coherent optics of objects in the range between dimers and nanoparticles.

Coherence may exist both in the internal and translational degrees of freedom. Internally, it may be preserved in the dynamics of electronic, vibrational, rotational and excitonic states [1214]. In contrast to that, our present review discusses the external translational (motional) evolution, that is pure de Broglie wave mechanics, which may be enriched by the particles' size, temperature and internal complexity: internalstate properties can eventually modify the motional dynamicsin the presence of external fields [15] or in scattering processes [16].

We aim to understand the evolution of many-body systems, which may comprise hundreds of atoms, and we are interested in effects that can be associated with the de Broglie wavelength λdB = h/mv, where m and $v=\vert \vec{v}\vert$ are the mass and centre-of-mass velocity of the composite particle, respectively, and h is Planck's constant. The following key questions are in the current focus of research:

If even very massive objects are allowed to propagate in compliance with the predictions of quantum wave dynamics, why don't we observe quantum delocalization in our macroscopic world? Will the quantum-classical 'boundary' simply shift with future advances of experimental refinements or is it reasonable to expect a fundamental limit, for instance imposed by a mass-scaling non-linearity in the underlying quantum equations [1719]? How well can we control decoherence [2022] to advance experimental quantum physics? Can we test new physics by exploring 'quantum macroscopicity' [2328]? Can gravity play a role at some point? Such questions are also relevant in the context of counter-propagating ring currents in superconducting quantum devices (SQUIDs [29]), in the coherent splitting of atomic Bose–Einstein condensates (BECs) [30] or in the superposition of micromechanical oscillator states [31]. Here we study the experimental basis for pursuing them in de Broglie experiments [32, 33], where we ask whether nanoparticle interference may allow us to search for modifications of established quantum mechanics [17, 34, 35].

Most of them rely on a non-linear extension of the Schrödinger equation and their spirit is always very similar: develop a master equation for the time evolution of the system's density matrix ρ, which will no longer be described by the conservative von Neumann term [Hρ]/i $\hslash$ alone. Instead, it is supposed to be enriched by a Lindblad superoperator L: ${\partial \rho}/{\partial t}=[H,\rho]/{i\hslash}+L\rho$ , where H is the Hamiltonian of the quantum system and the Lindblad operator may encompass standard decoherence [21], a spontaneous collapse of the wave function [17, 19], interactions with the gravitational field [36] or modifications of space–time [18, 37]. The principal goal of many interference experiments with massive particles is then to explore and quantify the relevance of this extension.

The devices that are being built to test such fundamental questions are also relevant for practical applications, such as quantum-assisted metrology [8, 15, 38]. Many matter-wave interferometers generate a free-flying periodic particle density pattern, which can be used as a ruler to measure tiny forces or molecular properties through interference fringe shifts in the presence of external fields [9, 10, 15, 39, 40]. Forces on the level of y N are easily measurable when molecules interact with external electric, magnetic or optical fields through their polarizability, magnetic susceptibility, absorption cross section or other internal properties.

The key idea behind these experiments is very analogous to that of atom interferometry: if we can divide a particle's wave function to travel along at least two spatially distinct paths and if we expose their distinct amplitudes to different force fields they will pick up a phase difference, which finally shows up as a modulation of the particle density and count rate behind the interferometer. In practice, the accumulated phase commonly grows with the enclosed interferometer area, both in space or in a space–time picture, and it often also depends on the mass of the particle. Since the interferometer area usually grows with decreasing forward velocity, many experiments dep end on the availability of slow particle beams. Their generation will be an important aspect of the following discussion.

Finally, we foresee a new avenue in the interference-assisted preparation and probing of nanostructures [41, 42, 194]. Molecules exhibit tiny de Broglie wavelengths (1–10 pm) at low kinetic energies (0.1–10 eV). Coherent matter-wave optics might thus allow us to deposit molecular nanostructures onto a surface [42] or to probe surfaces with molecules.

In section 2, we briefly review the history of molecular matter-wave experiments. Section 3 outlines methods to prepare neutral matter beams. We briefly discuss the concept of coherence and discuss techniques to increase the degree of coherence of molecular beams and we analyse different particle detection schemes in section 4. While section 5 is dedicated to the diffraction of molecules, section 6 discusses molecule interferometers that combine several optical elements to more refined instruments linked to the names of Ramsey and Bordé interferometer, Mach–Zehnder interferometer (MZI) or variants of Talbot–Lau interferometer. We finally distinguish molecular phase averaging and decoherence in section 7 and we give a short outlook in section 8.

In order to increase the legibility of the paper we have introduced an icon system (see table 1), which represents the building blocks of each experiment: some experiment can be conceptually derived from another by exchanging one or a few components. We also note that throughout the text we occasionally use the word 'molecule' as a placeholder for either 'cluster', 'molecule' or 'nanoparticle' when the particular character of the internal atomic bonds or arrangement is of no further relevance.

Table 1. Icons to guide the reader through the sections and experimental building blocks of molecular matter-wave interferometry.

Icon Description Icon Description
Thermal sublimation, section 3.1.1 Fluorescence detection in free flight, section 4.1.7
Supersonic beam, section 3.1.2 Scanning tunnelling microscopy section 4.2.1
Laser evaporation, p 6 Fluorescence microscopy, section 4.2.2
Neutralization of molecular ions, p 6 Two-dimensional diffraction mask, section 5.1
Slotted disc velocity selector, section 3.2.1 Diffraction at a disc/sphere, section 5.2
Helical velocity selector, section 3.2.1 Diffraction at a crystal, section 5.3
Gravitational velocity selection, section 3.2.2 Blazed material grating, section 5.4
Electron impact ionization, section 4.1.1 Material grating, section 5.5
Langmuir–Taylor ionization, section 4.1.2 Optical grating, section 5.6
Thermal photoionization, section 4.1.4 Ramsey–Bordé interferometer, section 6.1
Field ionization, section 4.1.5 Mach–Zehnder interferometer, section 6.2
Magnetic sector field, p 8 Three-grating Talbot–Lau interferometer, section 6.3
Two-dimensional multi-channel plate array, p 8 Two-grating Talbot–Lau interferometer, section 6.3.2

2. A brief history of matter-wave physics with molecules, clusters and nanoparticles

The origins of molecule interferometry can be traced back to as early as the year 1930 when Estermann and Stern demonstrated coherent scattering of H2 molecules at a clean LiF and NaCl surface [43] (see figure 1). The de Broglie wavelength of supersonically expanded hydrogen dimers at a velocity of v = 2000 m s−1 amounts to $\lambda_{{\rm dB}} = 1\,{\AA}$ . This is well compatible with the interatomic spacing of a clean crystal surface and the H2 molecules scatter with an angular distribution as explained by quantum diffraction.

Figure 1.

Figure 1. Experiments by Estermann and Stern in 1930 demonstrated quantum diffraction of helium atoms and molecular hydrogen at clean LiF and NaCl (shown here) surfaces. Following the pioneering experiments by Davisson and Germer [44] and Thomson [45], they were the first to corroborate de Broglie's quantum wave hypothesis [46] for composite and significantly more massive systems than electrons. The line is drawn as a guide to the eye.

Standard image High-resolution image

Coherent surface scattering has recently been revived with dimers at kinetic energies up to 10 keV and de Broglie wavelengths down to 100 fm [16, 47].

Even earlier than that, a number of molecule interferometers have been developed: Already in 1981, Christian Bordé and colleagues performed optical Ramsey spectroscopy on SF6 [48] but it took the scientific community about a decade to realize [49] that the internal state labelling provided by photon absorption is also accompanied by coherent momentum transfer. Various diffraction and interference experiments were performed with nanofabricated gratings: the weakly bound helium dimer He2 was discovered because of its diffraction pattern [50], Mach–Zehnder interferometry allowed us to explore collisional properties of Na2 [51] and diffraction of internally hot many-body systems was first realized with the fullerenes C60 and C70 [32, 52].

Dimers formed from Bose–Einstein condensed (BEC) are the most spatially extended molecules, by far. They may extend to dozens of micrometres close to their dissociation threshold [53]. Nowadays, BECs provide the coldest possible sources for quantum experiments with ground-state dimers [54] or with ultracold cesium tetramers [55].

A number of de Broglie experiments relied on supersonic molecular beams: magic clusters were revealed in the diffraction of helium clusters up to He100 [56, 57] and D2 was quantum imaged into the shadow behind a microdisc [58], thus implementing the idea of a Poisson spot with molecules [59].

Macromolecule interferometry has recently been extended in various directions: diffraction at a grating [41, 60] is conceptually close to the textbook example of a double slit, but it also requires a molecular beam, which is sufficiently coherent to cover at least two neighbouring slits by wavelets with a well-defined phase relation. This triggered efforts in Vienna and Southampton to implement a multiplexing near-field interferometer, which was first realized for atoms by Clauser [61, 62] and then implemented with molecules in Vienna—meanwhile in three different variants, namely Talbot–Lau interferometry (TLI [63]), Kapitza–Dirac–Talbot–Lau interferometry (KDTLI [64, 65]) and the optical time-domain ionizing matter-wave interferometer (OTIMA [6668]).

The TLI design requires typically three gratings for coherence preparation, for diffraction and to scan the interference fringe pattern. It was used for the first de Broglie experiments with biomolecules [69], to study collisional [70] and thermal [22] decoherence and to establish quantum-enhanced molecule metrology [71] and lithography [42].

The idea of a KDTL interferometer builds on the Talbot–Lau concept but the central grating is replaced by an optical phase mask to eliminate the attractive van der Waals force that arises during the interaction between matter and material diffraction gratings. The KDTLI concept enabled a series of interference experiments with the most massive particles to date [72, 73] and with new applications in quantum-assisted metrology [1, 8, 15].

The OTIMA concept [66, 67] extends this idea by replacing all mechanical gratings by optical standing light waves whose photon energy suffices to ionize or 'deplete' the incident molecular beam in the antinodes of the standing wave. Because of the tiny immaterial grating structure, this concept is believed to be particularly versatile and useful for future high-mass interference studies.

Several alternative schemes have been suggested for coherence experiments with nanoparticles [7476]. They all assume the availability of cooling schemes, which would allow one to trap and confine a particle's wave function so tightly that its high momentum uncertainty ensures a fast expansion of the wave function, once the particle is released.

In order to even advance the field of nanoparticle quantum optics, a number of challenges have to be overcome: new sources shall prepare cold and coherent molecular beams. New diffraction elements must be compatible with the rich internal structure of complex molecules. Decoherence has to be avoided and new detectors should combine high efficiency with position accuracy and selectivity to mass, chemical composition or internal states. In the following we document progress along the way and future perspectives.

3. Molecular beam sources and coherence

An analogy between electromagnetic and quantum wave physics can be established based on the fact that the stationary Schrödinger equation is formally equivalent to a Helmholtz equation, as it is found in the classical electrodynamics of scalar fields [77]. Also the concept of coherence can therefore be transferred to matter waves. We here divide the source requirements into those referring to the longitudinal and transverse coherence, respectively. This division may be questioned for ultracold particles with an isotropic velocity distribution, but it is justified in de Broglie experiments with collimated molecular beams, with vtrans/vlong = 10−3–10−6.

The degree of coherence [4, 77] quantifies the correlations and phase relations in a wave field. Two wavelets may interfere if they originate from the same cell in phase space and if they are fundamentally indistinguishable. For massive particles this includes their internal and external dynamics—from the source to the detector.

All molecular de Broglie experiments to date have realized single-particle interference, where partial wave functions of one and the same particle interfere and all molecules are prepared in similar—but not necessarily identical—states to add to the same fringe pattern. Two-particle interference would require the wave functions of two independent molecules to overlap in space while the molecules are equal in their atomic, isotopic and conformational composition. They should occupy the same electronic, rotational, vibrational and spin states. At present this can only be done for diatomic ground-state molecules in BECs. But experiments with these ensembles have remained in the regime of single-particle interference, see also [53, 78].

Longitudinal and transverse coherence are measures for the width of the Fourier transform both of the de Broglie wave spectrum and of the effective source emission function [79]. Their numerical value varies in the literature by factors of π, depending on whether the initial state is better described, for instance, by a Gaussian wave packet or a plane wave with finite support [77, 80].

As a rule of thumb, longitudinal coherence can be estimated by $l_{{\rm c}}\simeq \lambda_{{\rm dB}}^{2}/{{\Delta}\lambda}_{{\rm dB}}\propto 1/{\Delta}v_{{\rm long}}$ . It measures the spectral purity of the source. A measure of the lateral correlations is given by the transverse coherence, which is estimated by Xc ≃ λdBLx, where L is the distance to the source and ΔλdB is the spread in the distribution of de Broglie wavelengths and Δx is the size of the source.

A rather comprehensive treatment of various theoretical methods in matter-wave physics is provided by recent reviews on atom interferometry [2], molecule interferometry [1], as well as treatments on relativistic phases [81], Feynman path integrals [82] and a phase-space approach based on Wigner functions [83].

3.1. Molecular beam sources for quantum optics experiments

The goal of this section is to describe methods that aim at generating neutral molecular beams of sufficient flux and coherence to be compatible with molecule interferometry. We will discuss both established techniques and conceivable alternatives.

3.1.1. Effusive beam sources.

Thermal effusive sources have served as workhorses in various interference experiments with complex molecules [84, 85]. Stable particles can be heated up to the point where they sublimate or evaporate in a free effusive regime, where their velocity distribution is determined by a Maxwell–Boltzmann law. This is strictly only the case when the collisional mean free path of the molecules is larger than the smallest dimension of the source opening [85] and for that reason, one finds a supersonic contribution in many practical realizations, also.

The volatilization temperature has to stay low enough to avoid particle fragmentation. This is a challenge for complex particles with large polarizabilities and therefore strong van der Waals interactions to neighbouring molecules—which increase the enthalpy of evaporation. Based on this insight it is now possible to alleviate this problem by chemical tailoring the molecular bonds, appropriately [84]. The attachment of perfluoroalkyl chains to organic molecules by Mayor and co-workers in Basel has proven to be an important and very successful method that paved the path to the realization of quantum interference of particles in excess of 10 000 amu [72, 73, 84, 86].

The most probable velocity in a thermal beam is $v=\sqrt {2k_{{\rm B}}T}/m$ . This corresponds to a de Broglie wavelength of λdB ≃ 5 × 10−12 m in the case of C60 at 900 K [32]. Current state-of-the-art matter-wave interferometers can cope with λdB ≃ 200 fm [72, 87], corresponding to a particle of m = 105 amu thermalized at room temperature or m ⩾ 106 amu at 10 K.

The potentially detrimental effect of heating on the integrity of fragile molecules is a particular problem for biomolecules. The thermal load may, however, be reduced by limiting the heating power to a local spot, from which the molecules are rapidly laser evaporated. This method induces substantially less thermal damage than a Knudsen cell and it proved useful in far-field diffraction of tailor-made phthalocyanine derivatives [41]. A laser beam can be focused to less than a micrometre, which provides the transverse coherence that is required for molecule interferometry.

3.1.2. Supersonic sources.

Molecule interferometry has also been performed with particles entrained in a supersonically expanding carrier gas [43, 49, 50, 58, 87, 88]. The adiabatic expansion provides translational cooling and improves the longitudinal coherence as well as the forward-directed particle flux. This process affects the temperature of all degrees of freedom including translation, rotation and vibration with a temperature ranking Ttrans < Trot < Tvib [89]. The finite energy level spacing in molecules is the reason why vibrational modes may be completely frozen in small molecules [85], while in massive particles cooling to the rotational ground state is a big challenge. Jet cooling of internal states is only effective for molecules up to a few dozen atoms. The final temperature is determined by the molecular heat capacity as well as the number of thermalizing collisions with the carrier gas.

Supersonic beams have a wide range of applications since they can be loaded with free gases [43, 49, 50, 58] or particles that are volatilized in laser desorption (LD) [9092] or thermal sources [51].

Adiabatic expansion comes, however, at the price of an overall acceleration to velocities in the range of 300 m s−1 for cryogenic nozzles and more than 1000 m s−1 for helium at room temperature [51]. This high velocity can be reduced in the lab-frame when the nozzle is placed on a backward-rotating wheel [93]. At velocities below 100 m s−1 [94] small molecules would be compatible with matter-wave interferometry.

3.1.3. Future alternatives.

Beams of large neutral clusters may be formed in aggregation sources, based on magnetron sputtering [95], laser ablation or thermal evaporation [96]. They generate neutral or singly charged particles of either sign and are prime candidates for quantum interference experiments with metal or semiconductor clusters. Aggregation sources may produce beams of internally cold particles but still require further slowing or cooling for high-mass interferometry.

Charged particles may be acceptable precursors for molecular interference experiments. Charge may provide a handle for beam preparation and cooling. It seems, however, advised to add a neutralization scheme and to perform all quantum delocalization studies with neutral particles to avoid decoherence and phase averaging in stray electromagnetic fields. Matrix-assisted laser desorption ionization [91, 97] (MALDI) and electrospray ionization [92] (ESI) could serve as promising beam sources for such experiments. This is particularly true for thermolabile biomolecules of almost any size [98].

In MALDI, molecules are initially isolated in and desorbed from an organic matrix. The matrix molecules act both as absorbers for the desorbing laser light pulse and as collision partners in the following expansion into the vacuum. They are responsible for cooling as well as for charge transfer with the analyte particles. The result is a mixed molecular beam composed of matrix molecules and isolated analyte particles travelling at 300–800 m s−1. The method produces singly charged molecules and a substantial fraction of neutrals [99]. MALDI has been shown to volatilize objects with masses in excess of 1010 amu [99, 100], even functional biomaterials, which may become relevant for far-future quantum experiments.

A number of matrix-free optical methods are conceivable and currently being explored in several laboratories. This includes laser-induced acoustic desorption, LIAD [101], laser-induced forward transfer, LIFT [102], or matrix-free LD [103]. They are all compatible with a high-vacuum environment.

In contrast to that, ESI experiments start at atmospheric pressure and require the subsequent transfer of the isolated molecules through differential pumping stages into the ultra-high-vacuum science chamber. The molecules are first dissolved in a carrier liquid and filled into a thin capillary. High voltage (1–2 kV) between the capillary and the vacuum inlet leads to the formation of a liquid cone which fractions when the electrical forces overcome the surface tension (Coulomb limit). Release of small droplets and a repetitive cycle of solvent evaporation and charge instabilities finally lead to the isolation of individual solvent-free molecules. This method is again particularly useful for labile biomolecules. However, since electrosprays are intrinsically based on Coulomb fission, even a single protein of m ⩽ 105 amu ends up loaded with dozens of elementary charges. Charge reduction, efficient transfer into the vacuum chamber and slowing of the beam remain the key challenges for quantum experiments based on this source [66].

3.2. Preparation of coherence: transverse and longitudinal selection and cooling

With no molecular analogues to atom lasers at hand [104, 105] and transverse laser cooling still being demanding [86], the preparation of transverse coherence in molecular beams often relies on restricting the size of the effective source. This is either done by reducing the width of the emitter or by placing mechanical collimators to define the beam.

The longitudinal coherence length $l_{{\rm c}} \propto \lambda_{{\rm dB}}^{2}/\vert \Delta \lambda_{{\rm dB}}\vert = \lambda_{{\rm dB}}\cdot v_{{\rm long}}/\vert \Delta v_{{\rm long}}\vert$ increases with a narrowing of the velocity spread. A typical thermal beam source emits particles with a velocity distribution whose full-width at half-maximum Δvlong(FWHM) reaches up to 60% of the forward velocity v0 [32, 52]. Since the de Broglie wavelength depends on the particle velocity the diffraction fringe pattern itself can be used to determine or select a certain velocity, if the mass is already known [41, 106]. More commonly, the time-of-flight (TOF) between two well-defined points provides velocity information. If the particle beam emerges from a pulsed adiabatic expansion [68] a time-resolving detector is the appropriate instrument.

3.2.1. Rotating mechanical masks.

A continuous molecular beam can be made compatible with TOF selection by encoding a time sequence on it. The beam can be chopped using a single rotating disc with a slit. Two discs are needed to define a travel time when both the source and the detector are operating in continuous mode. In practice several, often up to six, discs are used to suppress higher order velocity bands [85]. The transmitted velocity range can typically be reduced down to Δvlong/v0 = 1%.

When a single disc is used in combination with a time-resolving detector, the signal can be enhanced by etching many slits into this disc. The resulting convolution of molecular packages travelling at different velocities and partially overtaking each other can complicate the velocity analysis. The signal can, however, be deconvoluted a posteriori if a known pseudo-random sequence is modulated onto the molecular beam [107]. Many different velocity classes can then be used in a single experiment.

The particle velocity can also be selected by a rotating turbine with helical grooves [42, 108]. Depending on the groove geometry and the rotational speed, only molecules within a certain velocity band can pass. This method has been widely used in neutron interferometry, but the vibrational noise of a mechanical device spinning at 100 Hz may also scramble quantum fringes: while far-field diffraction is rather inert against such perturbations [32, 51, 52] because the fringe spacing exceeds the grating period by orders of magnitude, vibration amplitudes as tiny as 10 nm may be detrimental for near-field interferometry [109]. It is, therefore, often advised to choose passive selection methods.

3.2.2. Vibration-free velocity selection by Earth's gravity.

In the early days of molecular beam physics, Otto Stern selected the molecular beam velocity by rotating the entire vacuum apparatus [110]. Nowadays, it appears rather favourable to employ a static scheme that exploits the free fall in the Earth's gravitational field [111]. Three horizontal slits placed between the source and the detector can define a free-flight parabola and select molecules that travel at a desired velocity [41, 111]. The selectivity of this scheme depends on the height and separation of all slits. It improves with decreasing velocity and enables the selection of velocity bands with a typical width of Δv/v ∼ 5 − 30%. In practice, the first and the third slit can be realized by the source and the detection aperture, respectively.

3.2.3. Slowing and cooling.

Slowing and cooling of complex molecules is still a grand challenge, but a number of sophisticated techniques have been demonstrated in recent years: laser cooling, which is tremendously successful for atoms [112], has recently been extended to diatomic molecules, such as SrF [113] and current efforts explore the possibility of going even further. It has also been shown that cold molecules could be generated by joining cold atoms [114, 115].

Since these techniques require cycling optical transitions and the exchange of many photons it is difficult to extrapolate them to compounds with dense and open energy level systems. A number of research teams have started sympathetic cooling of molecules co-trapped with laser-cooled atomic ions. This is promising but in order to trap two species in the same Paul trap it seems that the atom and the molecule should have the same mass-to-charge ratio [116120].

When optical cycling is no option, one may also exploit selective removal from the ensemble following the idea of Maxwell's demon: open and close a door between two chambers when you 'know' a slow particle wants to pass. An atom or molecule oscillating in any kind of trap reaches its lowest velocity at its turning point. When it is extracted from the ensemble at this very point, its motional energy is minimized. This trick has been implemented with atoms and it will be interesting to see how to extend it to larger objects [121].

It has been predicted that laser cooling of complex dielectric nanoparticles could also be realized using off-resonant optical forces in a cavity-assisted Sisyphus cooling scheme [122, 123]. Although several groups succeeded in proving the working principle for atoms [124, 125], substantial cooling of mesoscopic particles is still a challenge [126128].

Another concept relies on non-radiative but modulated electric or magnetic fields [129]. Since neutral molecules couple only weakly to external fields, many interaction steps are required to achieve sizeable deceleration. Electric [130132] and optical Stark deceleration [133], reverse magnetic coil guns [134, 135] and Zeeman slowers [136] have been used to manipulate molecules up to about 100 amu. Extrapolation of these methods to more complex objects is a subject for future research.

For gaseous polar molecules, Stark selection [137] rather than deceleration was shown to provide high phase-space densities and a good starting point for further storage and advanced laser cooling schemes [138140]. This technique works best when it can select from an existing dense cold and gaseous ensemble—which makes it less compatible with massive objects such as biomolecules or metal clusters.

Finally, cryogenic buffer gas cooling [141] and the implantation of molecules into liquid helium droplets [142, 143] are techniques that allow one to cool both the internal and the translational degrees of freedom. As many other schemes, they are still waiting to be integrated into preparatory stages for future matter-wave experiments with nanoparticles.

4. Molecular beam detectors

Molecular beam detectors serve various purposes [85, 144]. Apart from measuring molecular flux or density they may be sensitive to the particle's mass, internal state, position or velocity.

4.1. Ionizing detectors

Since ions can be manipulated in electromagnetic fields and counted and mass selected with high efficiency many matter-wave detectors comprise an ionization, mass selection and ion counting unit. We first focus on a discussion of different ionization strategies. Mass selection is then typically implemented through one of several well-established schemes, for instance using radiofrequency quadrupole guides (QMS), TOF spectrometers or magnetic sector field mass spectroscopy (SF-MS) [145].

While post-ionization of metals is generally efficient even at low energies, the ionization of noble gas clusters or large organic molecules turns out to be notoriously difficult [146148]: the ionization energies of most biomolecules range between 8 and 12 eV, i.e. a regime that is not accessible to commercial lasers. Interestingly, even dedicated extreme UV and x-ray light sources realized by plasma lamps or electron synchrotrons have not yet produced the expected breakthrough in biomolecular ionization mostly because of competing processes such as electron recapture, re-neutralization and fragmentation processes.

This may be circumvented by functionalizing (tagging) large molecules with smaller ones of low ionization potential [149] or by forming large clusters from well-ionizing smaller organic compounds [150]. A general method for post-ionizing or detecting complex bionanomatter has still to be developed. This is also the reason why it has been possible to demonstrate quantum delocalization for functionalized large molecules [72] or complex clusters of small organic dyes [87] in the mass range of small proteins, while quantum experiments with genuine proteins, DNA or RNA are still at the core of an intense open research programme. In principle, ion detection can be position sensitive with micrometre resolution, for instance when using multi-channel plates (MCPs).

4.1.1. Electron impact ionization.

When a molecular beam is bombarded with electrons of energy E the ionization efficiency ηion ∝ ne · σion(E) · l depends on the electron density ne, the molecular ionization cross section σion(E) and the interaction length l. For many molecules σion(E) is maximized for electron energies between 40 and 100 eV [85, 86]. The maximal electron density is limited by the energy-dependent space charge, i.e. by Coulomb repulsion between the electrons. With ionization cross sections ranging between 10−16 and 10−14 cm2, a total ionization yield of 10−3–10−5 can typically be achieved. For complex molecules, fragmentation is an important issue, since a large manifold of internal states enables fast conversion between electronic, vibrational and rotational energies.

In spite of these difficulties, electron impact was the method of choice for a large number of quantum diffraction and interference experiments with diatomic and organic molecules up to 104 amu [8, 15, 16, 39, 50, 58, 65, 69, 73].

4.1.2. Hot wires and Langmuir–Taylor ionization.

A neutral particle may also ionize when it hits a surface whose work function is higher than its own ionization potential. If the surface is sufficiently hot, the ionized particle may re-desorb and be collected by an ion counter [151]. That is the basic idea of Langmuir–Taylor or hot-wire detectors that have been widely used in alkali atom interferometers [51, 152].

For certain atom/surface combinations this process is highly efficient [153]. It is, however, less compatible with TOF mass spectrometry, since time constants between 10−6 and 10−2 s spoil the mass resolution [85]. On materials with a low work function and particles with a high electron affinity the reverse process, electron attachment, can also occur [154].

Some molecule/surface combinations, in particular those involving molecules, may favour molecular dissociation into products with lower ionization potentials [144]. In selected cases, up to the size of insulin, intact surface ionization has been reported in interaction with hot rhenium oxide, which exhibits a particularly high work function. Even then it was necessary to compensate for the missing potential energy by the kinetic energy of the incident hyperthermal molecular beam [155]. Since 'hyperthermal' is tantamount to saying 'very fast' this technique is not well adapted to macromolecule quantum experiments with long de Broglie wavelengths.

4.1.3. Single-, two- and multi-photon ionization.

The typical photoionization cross-sections of clusters and molecules up to 10 000 amu vary between 10−18 and 10−16 cm2 [149], i.e. about one hundred times smaller than that for electron impact ionization. In spite of that, photoionization is usually substantially more efficient. The reduced cross section can be overcompensated by the available photon flux, and photoionization is generally softer since it deposits less excess energy in the ionized particles.

Single-photon ionization (SPI) can be applied to organic molecules and clusters even above 10 000 amu [103, 150] and it has also been proposed [66, 67] and implemented [87] as the basis for a universal measurement-induced diffraction grating for matter waves.

If the energy of a single photon is not enough to overcome the ionization potential, a resonant or non-resonant two- (2PI) or multi-photon (MUPI) process may be used. Resonant MUPI can succeed if subsequent absorption events occur faster than the competing internal relaxation processes. In that case, the absorbed energy just barely exceeds the ionization potential and MUPI can also be a soft ionization scheme [156158].

4.1.4. Delayed ionization of laser-heated molecules.

Multi-photon thermal ionization has been studied for highly stable molecules, such as the fullerenes C60 and C70 [159, 160]. When they are internally heated to energies between 50 and 200 eV several competing processes may be observed [161, 162]: the emission of thermal radiation, the emission of electrons and finally fragmentation.

The rotational and vibrational modes may couple energy to the electronic states and, given the thermo-statistical nature of the energy randomization, ionization may be delayed by many microseconds after the exciting laser pulse.

Delayed ionization was an enabling technique for the first fullerene diffraction experiments [32] where the total ionization efficiency was estimated to be above 10% and the spatial resolution was fixed to a few micrometres, as determined by the waist of the laser beam [163].

4.1.5. Field ionization.

When a positive electric potential is applied to a tip as sharp as 10 nm, it is possible to create a local electric field of up to 1010 V m−1. A molecule in the vicinity of the needle will be polarized and attracted to the tip. The internal states are then shifted to the point that an electron can tunnel between the molecule and the surface. The molecular ions can again be counted [164]. Since field ionization is a soft process it is also applicable to complex molecules [144]. However, the effective interaction region around the tip is limited to 200 nm [85], which is both an asset and a drawback: it can be used to image a molecular beam distribution with high resolution, but it also takes a long time to scan a two-dimensional (2D) distribution. When the interaction zone is reduced to the size of a single atom [165] the emitter also becomes promising as a coherent source for ion interferometry [3].

Field ionization may also be interesting for studying Rydberg molecules, since the required electric fields can be very low for such highly excited molecules [136]. The combination of molecular Rydberg physics with matter-wave interferometry is still open for future research.

4.1.6. Ion counting techniques.

For particles with masses up to about 104 amu the overall detection efficiency is dominated by the ionization process. At higher masses the subsequent secondary electron yield ηSE becomes an additional factor. Depending on the particle species and the overall energy, ηSE ∝ vn grows with a power law in the ion velocity v, where n = 4 is typically valid for velocities up to 104 m s−1 [166]. Efficient detection, therefore, requires acceleration voltages of several 10 kV for masses beyond 105 amu.

While commercial mass spectrometers are equipped with secondary electron multipliers to register energetic ions, high-mass ions may also be counted using image charge detectors. Their current sensitivity of 1–10 fA sets a detection limit of about 104 elementary charges per detector bin and second.

4.1.7. Optical detection in free flight.

Small and cold molecules with narrow optical resonances and strong absorption lines are routinely detected via optical ionization, absorption or fluorescence in free flight. The access to specific internal states allows one to couple and even entangle internal and external degrees of freedom. This has been exploited in molecular Ramsey Bordé interferometry [48].

Free-flight optical detection is, however, hampered for complex molecules since they lack closed fluorescent transitions. Only in selected cases the short light–matter interaction time will be sufficient to capture sufficient fluorescence to detect a single particle. Also, the repetitive deposition of Stokes energy in the particle can lead to its rapid heating and damage. Such limitations are overcome when the molecule is bound to a solid substrate. Fluorescence can be collected over extended times and the substrate stabilizes the molecule by dissipating the Stokes energy.

4.2. Surface-based detection schemes

Surface-deposited clusters and molecules may be visualized by several modern microscopy techniques, ranging from scanning tunnelling (STM) and atomic microscopy (AFM) to near-field scanning optical microscopy (NSOM) or fluorescence microscopy in one of its many modern high-resolution implementations.

Depending on the method, the molecules can be identified by their size, shape and tunnelling spectra in tunnelling microscopy or by their absorption and emission bands in fluorescence microscopy. For this to be useful, it is important that the particles of interest are sufficiently distinct from all other species in the beam and on the substrate. Furthermore, it is crucial that surface diffusion processes [167] are negligible to prevent the deposited molecular interferograms from washing out.

Probing a surface-deposited quantum interferogram combines high detection efficiency with high spatial resolution. Moreover, intensity fluctuations of the molecular beam do not affect the measurement outcome since the entire fringe system is recorded simultaneously. A number of systematic uncertainties are therefore eliminated. This is to be balanced against the difficulty to monitor grating drifts or interferometer phase shifts in real time.

4.2.1. Scanning tunnelling microscopy.

Scanning tunnelling microscopy is known to excel in imaging single atoms and molecules [168171]. Given its high spatial resolution and the possibility of post-processing surface-deposited interference patterns, this technique is also suitable for exploring quantum-assisted nanolithography.

In molecule interferometry it has been applied as a fullerene detector, specifically for C60 molecules. Figure 2 shows an atomically resolved, flat silicon 7 × 7 (1 1 1) surface reconstruction, which provides 19 dangling bonds per unit cell and therefore plenty of possibilities for each fullerene to bind in the vicinity of where it lands on the surface, even at room temperature [172]. The surface physics involved in this process has been studied in detail by others with an interest in doped fullerenes as quantum logic elements [173]. The specific preparation of a reconstructed silicon surface requires ultra-high vacuum conditions around 10−10 mbar. This is a challenge in combination with molecular beams that may even be initially created at the mbar pressure level.

Figure 2.

Figure 2. Isolated C60 molecules deposited and immobilized on an atomically resolved Si (1 1 1) 7 × 7 surface. Even some intra-molecular features can be identified (see the inset) [42]. Copyright 2013 by the American Physical Society.

Standard image High-resolution image

STM imaging offers the highest possible spatial resolution that is physically meaningful for the purpose of macromolecule interferometry, but a single image scan may take 45 min for an area as small as 2 µm2.

4.2.2. Fluorescence microscopy.

Fluorescence microscopy provides an excellent tool to discriminate molecules of known absorption and emission bands from their surroundings. Single-molecule fluorescence imaging was introduced about 20 years ago [174, 175] and it has found numerous applications in physics, chemistry and biology ever since. Recently, it has been used to record the build-up of a quantum interference pattern from individually arriving phthalocyanine molecules in real time (see figure 3) [41].

Figure 3.

Figure 3. Single-molecule fluorescence detection of surface-deposited phthalocyanines. The graph shows the time evolution of the fluorescence intensity of two molecules attached to the quartz surface with a separation of about 5 µm. This distance can be easily resolved, optically. Constant emission followed by an abrupt disappearance of the entire signal is indicative of true single-molecule fluorescence. Each molecular peak can be fitted by an Airy function whose centre can be determined to much better than Abbé's resolution limit, here to about 10 nm. Picture reproduced from [41].

Standard image High-resolution image

Fluorescence microscopy [176, 177] can be directly interfaced to matter-wave interferometry when the molecular fringe pattern is deposited on a quartz window that closes the vacuum chamber. All optics can be mounted in air.

One may object that the spatial resolution of tunnelling microscopy exceeds that of optical microscopy by two or three orders of magnitude, because of Abbé's diffraction limit for optical imaging devices. In recent years, however, several methods have been shown to overcome these limits in optical microscopy [178183]. They rely on the selective isolation of molecules by bleaching or on non-linear optical processes and they might eventually enable the detection of molecular near-field interference patterns with features much smaller than the wavelength of light, down to the level of a few nanometres.

With a fixed optical resolution, the position accuracy in the measurement of each individual molecule can be enormously increased when the molecular surface density is low—which is even desired in single-particle interferometry [184]. If two neighbouring molecules are further apart than the width of the point spread function (PSF) of the imaging system, one can fit the PSF to the detected signal for each molecule to determine its centre. The ultimate accuracy is determined by the signal-to-noise ratio in the image. Values as tiny as 1.5 nm [184] and 10 nm [41] have been achieved in recent experiments.

Optical microscopy hinges on molecular parameters as well as the experimental design. Dye molecules, such as for instance phthalocyanines and rhodamines are particularly well suited for that purpose, since they combine high optical absorption cross-sections and a high fluorescence quantum yield with long bleaching times. Larger molecules can be tailor-made and this enabled far-field diffraction with the most massive molecules to date.

Similarly, almost all proteins and other biomolecules can be labelled with a wide range of dye markers, if they are not already fluorescent, such as for instance the green fluorescent protein GFP [185]. Many other nanoparticles, in particular quantum dots, also show strong fluorescence.

Recent studies have shown that even non-fluorescent nanoparticles can be optically detected either by their stimulated emission [186], their absorption [187189] or a change in the refractive index [191]. In order to reach a sufficient signal-to-noise ratio, these methods need a tightly focused and scanning probe beam, which increases the recording time when a wide area needs to be imaged.

4.2.3. Cryogenic and superconducting bolometers.

Neutral molecules could also be detected by cryogenic bolometers. While these are not sensitive to the molecular mass, they are sensitive to energy. This can be used in experiments, where the final internal energy is correlated with the de Broglie phase accumulated throughout the interferometer. The detection principle would be based on the transfer of energy from internal excitations to the detector. When the detector is cooled to its superconducting state the transferred energy can trigger the phase transition to the normal conducting state, which is detected as an increase in the electronic resistance. Already in the early days of molecular coherence, with SF6 molecules in a Ramsey–Bordé interferometer, the output was monitored by recording the energy increase that resulted from a resonant laser excitation of a specific vibrational state in one arm of the interferometer [48]. In future experiments, this idea may be extended to a large class of molecules when it becomes possible to refine the detector sensitivity to the level of single particles. This is expected to be achievable with superconducting single-particle detectors (SSPD) that have emerged over recent years [190].

A key challenge here is to keep the detector sufficiently clean throughout every experiment to avoid energy dissipation in a cushion of surface adsorbates [192].

5. Quantum diffraction of molecules

With a bundle of molecular beam generation and detection methods now at hand, the following section is dedicated to coherent beam splitting methods that have already been realized with molecules. More complex compositions of these matter-wave optical elements to full-fledged interferometers and their applications will then be discussed in section 6. For navigation purposes, the figures in this section are complemented with icons that link to the sections that describe the relevant experimental techniques.

5.1. Molecular speckle patterns: diffraction at arbitrarily shaped pinholes

An arbitrarily shaped pinhole is probably the simplest and most generic of all possible diffractive objects. It has been studied using a supersonic beam of neutral H2.

The Fourier image of a micrometre-sized arbitrarily shaped mechanical aperture leads to a speckle pattern (see figure 4) [193], which can be regarded as a first step towards diffractive coherent imaging of microscopic objects [194]. Field ionization at a tungsten tip served the detection of H2 very well. The limited detection area required scanning of the tip but 2D field ionization arrays may become part of future wide-area detectors [164].

Figure 4.

Figure 4. Diffraction of H2 at an arbitrarily shaped aperture. (a) Scanning electron micrograph of the hole, and (c) the observed interference pattern. (b) Theoretically expected speckle pattern. The figure is adapted from [193]. Copyright 2006 by the American Physical Society. The pictograms introduced here on the right-hand side summarize in a single view the source (top), diffraction method (middle) and detection mechanisms (lower panel/s).

Standard image High-resolution image

5.2. Poisson's spot: diffraction at an opaque obstacle

An opaque object of cylindrical symmetry is the inverse structure to the pinhole discussed before. Wave theory predicts a bright spot at the centre of the shadow region behind the object, when it is illuminated coherently. Historically, an experiment performed by Arago and named after Poisson was crucial to defend Fresnel's wave theory of light. Recently, such an experiment has also been successfully performed with D2 matter waves [58] (see figure 5).

Figure 5.

Figure 5. Poisson spot seen with D2 diffracted at an opaque circular disc [58]. This is the molecular analogue to an earlier experiment that showed diffraction of atoms around a wire [197]. Picture reproduced from [58]. Copyright 2009 by the American Physical Society.

Standard image High-resolution image

At first sight, the Poisson experiment appears very promising for probing the wave nature of large molecules as well: due to the symmetry of the arrangement, alignment is easy and the Poisson spot is 'white', i.e. all wavelengths interfere constructively on the optical axis behind the obstacle. Since the bright spot appears at the centre of the classical shadow region, one might even think that it is background-free.

It is, however, important to see that the van der Waals force between the molecule and the obstacle also classically attracts all particles towards the shadow centre [195]. Adding to that a realistic velocity distribution, finite edge effects, the finite source size and a limited detector resolution one finds that the distinction of classical and quantum shadows becomes increasingly difficult for increasingly massive objects. For molecules in the size range of fullerenes the experimental prospects are still promising [196].

5.3. Diffraction of fast molecules at crystal surfaces

In contrast to the two previous experiments, which probed microscopic structures in transmission, a number of experiments focused on probing nanostructures in molecular reflection. Coherent scattering of molecular beams can thus be used as an analytical tool to study the potential energy surface of crystals [198]. These experiments are preferentially performed with light molecules, such as H2 or D2, at energies of about 10–100 meV, corresponding to de Broglie wavelengths between 50 and 200 pm [16].

Recent studies have extended this theme to fast molecular beams with energies up to 25 keV and de Broglie wavelengths down to λdB ≃ 100 fm [199, 200]. The molecular beam was directed at a crystal surface under grazing incidence to observe quantum diffraction and channelling within the surface potential. A typical interference pattern is shown in figure 6.

Figure 6.

Figure 6. 2D intensity distribution for scattering of H2 at E = 0.6 keV under grazing incidence at a sulfur superstructure on Fe(1 1 0). Molecular intensity increases from blue to red. Reprinted from [201], with permission from IOP Publishing.

Standard image High-resolution image

5.4. Reflective quantum diffraction at micromechanical gratings

Molecule diffraction can act as a surface probe as before, but surface scattering can also reveal specific characteristics of molecules and textbook examples of quantum physics. Quantum mechanics teaches us that, contrary to the situation in classical physics, a particle may be reflected by an attractive potential well. This has recently been shown for He2 and He3 reflected by a blazed diffraction grating [202]. When these weakly bound molecules approach the grating under grazing incidence they approach a van der Waals potential that is four orders of magnitude bigger than their internal binding energy (∼10−7 eV). The fact that molecular dimers with a mean bond length of 5.2 nm are not torn apart by the external potential shows that quantum reflection must take place several tens of nanometres above the surface. Under grazing incidence the molecular wave vector normal to the surface is sufficiently small to realize diffraction even at a grating with a period as large as d = 20 µm, which is orders of magnitude wider than the size of the molecular de Broglie wavelength (see figure 7).

Figure 7.

Figure 7. Quantum reflection of He2 and He3 at a blazed diffraction grating under grazing incidence. The angles, at which the nth order diffraction peaks of He2 and He3 are expected, are indicated by the red and dashed blue lines, respectively. Picture adapted from [202]. Reprinted with permission from AAAS.

Standard image High-resolution image

5.5. Transmission line gratings

5.5.1. Manufacturing of transmission gratings.

Since quantum reflection is a peculiarity of small and less-polarizable particles, transmission line gratings have been regarded as more generic and essential in a large number of studies with molecules. SiNx gratings are technologically particularly interesting since they remain pre-stressed and stretched even when they are pierced with holes of 50–100 nm width in periods of 100–300 nm [32, 42, 50, 51]. Far-field diffraction at a single grating is tolerant to small deviations from perfect periodicity and the required nanomask can readily be written using focused ion beam machines [41]. The concatenation of three gratings to a full interferometer requires additionally extraordinary reproducibility and precision between independent masks. This is especially important for wide-beam Talbot–Lau interferometers [51], where the average periodicity error needs to be smaller than 1 Å [65, 83]. Photolithography can provide large masks with a precision of Δd/d ≃ 10−5 in the grating period. This feat can actually even be achieved for gratings with a period down to 100 nm, using achromatic interference lithography [203].

5.5.2. General aspects of grating diffraction: near-field versus far-field.

Textbooks teach us that far-field interference peaks appear at sin βn = nλdB/d, where β is the angle to the optical axis. Near-field physics is usually less well covered in books but equally important for a number of coherence experiments.

In order to visualize the transition between both regimes we plot the evolution of a plane wave diffracted at a grating in figure 8. The vertical axis represents the decadic logarithm of the scaled distance behind the gating, where L/LT is measured in units of the Talbot length LT ≡ d2/λdB. The horizontal axis is linear in the scaled position $x/\bar{x}$ , where $\bar{x} = \lambda_{\rm dB}L/d +N\lambda_{\rm dB}L_{\rm T}/{6\rmd}$ . In the far-field (L ≫ LT) the scaled position corresponds to the diffraction angle in units of β1, i.e. the angle to the first diffraction maximum. In the near-field, i.e. for L < LT, $\bar{x}$ reduces to Nd/6L, which is proportional to the width of the grating.

Figure 8.

Figure 8. Evolution of the wave intensity behind a diffraction grating, from the near-field (top) to the far-field (bottom): The grating is coherently illuminated by a plane wave. Close to the grating, in the near-field, shifted self-images of the grating occur at multiples of the Talbot distance. The logarithmic distance scale allows us to see the full evolution in one picture. The lateral position scaling is described in the text.

Standard image High-resolution image

The scaling in figure 8 allows us to follow the wave evolution from the emergence of the Talbot carpets in the near-field (top) [204, 205] to the domain of far-field diffraction with well-resolved side peaks (bottom).

5.5.3. Applications of transmission line gratings: quantum-assisted mass spectrometry.

An interesting application of de Broglie interference in physical chemistry is to prove the existence of the weakly bound He2 through a mass assignment of its molecular diffraction peaks [50].

The helium dimer usually escapes mass spectrometry as it is bound with an energy of about 1.1 mK (100 neV) [206], which is too small to survive impact ionization with typical electron energies in the range 60–100 eV. The de Broglie relation, however, allows one to deduce the mass of a particle from its diffraction angle at a given nanostructure when its velocity is known (figure 9). Supersonic expansion of helium from a small nozzle leads to a fast and well-collimated beam, which is sufficiently cold and dense to form helium dimers and clusters. Transverse coherence can be established by tight collimation, longitudinal coherence by the intrinsic velocity compression during the adiabatic expansion of the noble gas jet. It is interesting to see that diffraction at a mechanical mask is sufficiently non-invasive to leave the dimers intact during the scattering process, in spite of the fact that the kinetic energy exceeds the intra-molecular binding energy by orders of magnitude. The particles are thus first sorted by quantum interference and then detected using electron ionization quadrupole mass spectrometry.

Figure 9.

Figure 9. The analysis of interference patterns allowed the detection of weakly bound clusters, up to He100 atoms. The different lines represent the cluster diffraction curves that appear for different source pressures P0. Larger clusters are formed at a higher stagnation pressure and show up under a smaller diffraction angle. Reprinted with permission from [57]. Copyright 2006, American Institute of Physics.

Standard image High-resolution image

Far-field diffraction was also used to observe the quantum behaviour of supersonically expanded Na2 [51] and later also of clusters up to about He100. This also facilitated, in particular, the investigation of 'magic clusters', i.e. clusters of particular structural stability [56, 207] and bound HeH2 halo molecules [208].

5.5.4. Applications of transmission line gratings: establishing quantum superposition at high mass, temperature and complexity.

All experiments described in this report are compatible with the rules of quantum mechanics. While this is a rather satisfying fact it is still in marked discrepancy with the observation that we do not see quantum superposition in our daily lives. Many reasons can be invoked for this phenomenon: the kinematic explanation of non-interference is inherent to quantum physics and points to the fact that the de Broglie wavelength of macroscopic objects is just way too small for us to have hope to ever resolve it. Equally well rooted in quantum mechanics are the predictions of decoherence theory [20, 21, 209212], which points to the fact that quantum observations are usually made on isolated systems. Once the individual system is coupled to a complex environment, quantum information diffuses into a large ensemble and becomes essentially unobservable in the partial subsystem. Genuine decoherence is based on quantum entanglement between systems. Quantum physics itself thus ensures that certain quantum phenomena become unobservable on a large scale. Also experiments with molecular matter waves corroborate the presence and significance of this effect [22, 70].

This also motivates a growing set of quantum interference experiments with massive and complex molecules, clusters and nanoparticles [1, 33], which was experimentally initiated with a demonstration of quantum interference with C60 and C70 [32] (see figure 10).

Figure 10.

Figure 10. Coherent diffraction of C60 molecules at a nanomechanical grating. Experiments with fullerenes C60 and C70 were the first to demonstrate quantum interferometry with thermally excited complex compounds. In spite of internal temperatures of the order of 900 K they exhibited high-contrast interference in their centre-of-mass motion. Data are fitted with a Fraunhofer diffraction equation taking into account an effective narrowing of the slits due to van der Waals forces between the molecules and the grating walls [32, 52]. Reprinted with permission from [52]. Copyright 2003, American Association of Physics Teachers.

Standard image High-resolution image

When teaching quantum interference to students we usually discuss the deterministic fringe pattern that is formed from individually and stochastically arriving quanta on a screen at some distance behind a diffraction grating. This aspect of the wave–particle duality has recently been visualized with molecules in two experiments, where an interference pattern was first deposited onto a substrate and then imaged using either fluorescence [41] or tunnelling [42] microscopy. They both offer single-molecule sensitivity and a spatial resolution down to the molecular level in two dimensions.

Figure 11 shows the quantum fringe pattern that was recorded with phthalocyanine molecules after diffraction at a SiNx mask with a period of 100 nm and a membrane thickness as small as 10 nm. Due to the gravitational velocity selection applied in these experiments, slower molecules are deposited further down on the detection screen. Their longer de Broglie wavelength implies larger diffraction angles and therefore larger separations between the interference orders than those observed for the fast molecules, which arrive further up on the screen. An area of 400 × 400 µm2 can be imaged within a few seconds, orders of magnitude faster than using STM or AFM techniques. A sequence of images recorded at different deposition times can show how the deterministic fringe pattern is built-up from individual and randomly arriving molecules.

Figure 11.

Figure 11. Single-molecule imaging in real time allows visualizing the wave–particle duality of fluorescent dyes; here of phthalocyanine [41]. Each dot represents a single molecule after its passage through a nanomechanical diffraction grating. The ensemble shows that the propagation of each particle must be described by quantum mechanics. The scale bar is 20 µm wide.

Standard image High-resolution image

5.5.5. Applications of transmission line gratings: molecular diffraction assesses van der Waals forces.

Material gratings are often regarded as binary transmission masks, i.e. as purely absorptive structures. However, due to the van der Waals interaction of each polarizable molecule with the grating bars we also have to consider an additional phase contribution [41, 83]. Quantum physics causes that every polarizable particle undergoes stochastic charge fluctuations, typically most pronounced on the femtosecond time scale. They induce image charges in near-by surfaces such as the side wall of a diffraction slit. The resulting attraction to the wall has three consequences: first, it removes particles if they get too close to the wall within a cut-off distance xc. Second, it dephases the matter wave, the stronger the closer the molecule approaches its image charge. Third, this phenomenon is also dispersive. Slow molecules acquire a stronger dephasing than fast ones. While a correct description of the van der Waals forces requires a detailed knowledge of the frequency spectrum both of the molecule and the surface, we estimate the cut-off distance based on the computationally simpler Casimir–Polder potential, VCP = −C4/r4. One finds xc = (18C4b2/mv2)1/6 to be of the order of 20 nm for gratings with a thickness of b = 100 nm and molecules with a static polarizability of the order of $\alpha \simeq 100\,{\AA}^{3}\times 4\pi \varepsilon_{0}$ , a typical mass of 500–700 amu and a velocity around 150 m s−1 [83]. The potential VCP generally scales with the fourth inverse power of the distance or length parameter r, while the interaction strength is defined by the constant C4. More information about the construction of the interaction constant can be found elsewhere [213].

These effects have been investigated in great detail in diffraction [214, 215] and Mach–Zehnder interference [216] of atoms and small molecules and they turned out to reduce the effective open slit width in the diffraction of fullerenes by 50%, for grating slits with an initial width of 50 nm [32, 52].

This indicates that the Casimir–Polder approximation overestimates the cut-off xc.

To minimize these effects, recent quantum diffraction experiments with phthalocyanines were performed with gratings of only 10 nm thickness. And yet the van der Waals forces are still substantial, as also seen in figure 11. Further experiments will, therefore, be needed to explore the extreme limit of material gratings, which shall finally even enable quantum diffraction at graphene, i.e. a membrane whose thickness measures only a single atom.

5.6. Optical manipulation of molecular matter waves

Optical gratings are very appealing for matter-wave experiments for various reasons. They can be realized with the high periodicity that is determined by the quality of modern lasers and mirrors. They can be timed, their open fraction can be changed both in situ and in real time, and optical phase gratings are also perfectly transparent for the particle beam. Optical absorption gratings can be realized without the influence of van der Waals dephasing.

5.6.1. Optical absorption and ionization gratings.

An optical absorption grating is based on the spatially periodic depletion of the molecular beam. It thus represents the idea of a measurement-induced grating [217] based on the selection of positions at which the molecules are removed from the beam—either by extracting them or by pumping them into non-detectable states [218].

Various molecules undergo SPI at photon energies beyond 7.9 eV, the energy of a VUV photon at 157 nm. One can achieve high ionization probability at the antinodes of a standing light wave and full transmission of neutral particles at its nodes. Since ions can be easily extracted by electric fields an optical ionization grating is effectively absorptive [66, 68].

Molecular absorption cross-sections between 10−18 and 10−16 cm2 necessitate the use of pulsed lasers with energies in the range of 1 mJ or very high-power (cavity-enhanced) continuous lasers. Pulsed lasers are advantageous as they also enable the realization of interferometers in the time domain [67, 68]. In practice, the molecules will not only experience a spatially modulated transmission but also a position-dependent phase shift, which is determined by the optical dipole force.

5.6.2. Dipole force and phase gratings.

The success of atom optics throughout the last decades was strongly driven by methods to manipulate the internal and external degrees of freedom with tailored laser light [219]. In contrast to atoms, with resonant absorption cross-sections as high as $\lambda_{{\rm dB}}^{2}/2\pi \simeq 10^{-9}\,{\rm cm}^{2}$ and strong optical polarizabilities close to narrow resonance lines, molecular line widths can extend to dozens of nanometres. Molecular absorption cross sections therefore typically range between 10−18 and 10−15 cm2.

Even through there is no substantial resonance enhancement of dipole forces in warm and complex molecules, recent experiments have successfully utilized the coherent interaction between a non-resonant standing light field with a molecular optical polarizability αopt: the optical field E induces an oscillating dipole moment, which interacts again with the optical field. The resulting force F = −αoptE2/2 depends on the intensity gradient.

In atomic physics this force can be attractive or repulsive, depending on the relative detuning between the laser frequency and the internal resonance line. While this choice is still possible in diatomic molecules, most complex compounds studied so far were effective high-field seekers, i.e. attracted towards the field maximum.

When laser light is retro-reflected at a smooth mirror surface it generates a standing wave with a periodicity of λL/2 whose periodicity and contrast depends on about half the coherence length of the incident laser beam. This may extend to thousands of kilometres for narrow-band continuous wave (cw) lasers and it is limited to several millimetres for pulsed excimer lasers, in practice.

When a spatially extended molecular matter wave falls onto a dipole phase grating, the incident molecules will leave the interaction zone in a superposition of transverse momenta with po = pin + Δp and ${\Delta}p=2{\rm n}\cdot \hslash k$ , where n ∈ N. The relative weight of the nth momentum states is determined by the Bessel function Jn(φ), where the modulation index φ ∝ PLτ/wxwy is proportional to the laser power PL, its beam waists wx and wy and the molecular transit time τ ∝ 1/vlong.

Since the molecular lines are broad and the interaction cross-sections are small, intense and well-focused laser beams are required to induce a significant dipole force. At the same time, it is important to avoid real absorption processes, which would be accompanied by three effects that are usually undesirable.

First, photon absorption can be followed by partial non-radiative internal conversion of its energy into excitations of vibrational and rotational states. A repeating sequence of many such processes may eventually heat the molecule to a degree that it fragments.

Second, photon absorption may be followed by spontaneous reemission and momentum recoil whose direction varies stochastically from event to event. Integration over many molecules and many recoils then results in a smearing of the interference pattern.

Third, it is interesting to realize that absorption from a coherent laser wave does not induce decoherence, by itself. Removal of a photon from a coherent beam does not leave any information behind and all phases are conserved. Pictorially, the interferogram is only laterally shifted by a fixed amount, which is related to the recoil of a single photon. Since each photon is in a superposition of two momentum states—being incident onto the mirror and being reflected—the absorption of a single photon is by itself a coherent beam splitting process with momentum transfer $\Delta p=\pm \hslash k$ (see [220] for single-photon emission close to a mirror).

The internal heating associated with the absorption and internal storage of a photon does not deteriorate the de Broglie interference pattern as long as it does not invoke any photo-activated or thermal emission process, which could release 'which-path' information to the environment. Even in spite of its coherent nature absorption will reduce the interference contrast, since the diffraction peaks related to the dipole phase grating are separated by $\Delta p=2n\cdot \hslash k$ and will be interlaced with those related to the single-photon absorption peaks at $\Delta p=\hslash k$ or even higher (odd) orders at high laser intensities. This relates molecular diffraction at optical gratings to studies of quantum random walks as well [221].

In [60] a green laser beam was reflected from a mirror to form a standing light wave with a period of 257 nm at which a coherent beam of fullerenes C60 was successfully diffracted, as shown in figure 12.

Figure 12.

Figure 12. Diffraction of C60 molecules at an optical phase grating. A collimated and coarsely velocity-selected (Δv/v ≃ 1/6) fullerene beam is phase-modulated by the interaction with a standing laser light wave. The far-field diffraction pattern is described by the Fourier transform of the optical potential. (a) Free molecular beam, (b)–(d) far-field diffraction pattern for increasing laser power. In all cases, the probability for a molecule to absorb a photon from the standing light wave is less than one. Diffraction at phase gratings allows the suppression of the forward-directed zeroth-order beam. Reprinted with permission from [60]. Copyright 2001 by the American Physical Society.

Standard image High-resolution image

The interaction time of molecules with a cw optical grating is velocity-dependent (dispersive) and the phase shift reads $\Delta\Phi \propto\alpha_{{\rm opt}}E^{2}(xyz)/v\hslash$ . Even though the spatial phase gradient is much weaker than that in the case of van der Waals interactions it is still necessary to pre-select a velocity class to maintain a high interference contrast.

The advantage of light gratings is particularly important in quantum interference experiments with highly polarizable molecules. Light gratings were therefore also integrated into KDTL [65, 222] and OTIMA [68] interferometry.

6. Advanced molecule interferometry with multiple optical elements

Far-field diffraction of massive matter is probably the most intuitive and pictorial realization of a quantum wave phenomenon. Yet, although it is still conceivable to extrapolate this method to nanoparticles in the mass range of 106 amu or more in a fountain configuration [195], more sophisticated experimental arrangements are desirable for certain applications, such as molecular precision metrology and nanoparticle interferometry.

6.1. Ramsey–Bordé interferometry measures molecular spectra and transition moments

The first molecule interferometer composed of several beam splitters was built with the purpose of high-resolution spectroscopy in SF6 [48, 49]. The introduction of separated oscillatory fields by Ramsey [223] had led to a boost in spectroscopic resolution and it was therefore translated into molecular physics. Such experiments start by creating a superposition of two internal states whose different phase evolution in free flight is finally converted into a detectable population difference. Figure 13 shows how this is implemented in practice: when a laser beam interacts with an effective atomic or molecular two-level system, it induces Rabi oscillations [224], i.e. periodic population oscillations between the two coupled states. The collimated molecular beam crosses the first running laser wave with an interaction time and laser intensity adjusted to realize a 'π/2-pulse' yielding a coherent and equal superposition of both states. In quantum information language this beam splitting corresponds to a Hadamard gate [225]. At the same time, this process generates entanglement as it couples the internal and external molecular degrees of freedom such that neither of them is known before a measurement, but both are inseparably coupled. If the atom stays in the ground state |g〉, the initial momentum also remains unchanged. If it gets excited to |e〉 it receives a recoil by the momentum of one laser photon $\hslash \vec{k}$ : $\psi \propto |g, \vec{p_{0}}\rangle + {\rm e}^{{\rm i}\phi}\vert e\vec{p_{0}}+\hslash \vec{k}\rangle$ . Stimulated emission by the second laser removes this momentum again and takes the molecule back into the ground state. A similar sequence of two laser pulses is then repeated to close the molecule interferometer, i.e. to recombine the two possible beam paths.

Figure 13.

Figure 13. Ramsey–Bordé interferometry with molecules: Ramsey's method of separated oscillatory fields was originally implemented for molecular beams with the goal to achieve a new level of high-resolution spectroscopy. Already in 1981 this technique was employed on molecules as complex as SF6 [48]. The beam splitting was, however, not yet spatially resolved because of the width of the incident molecular beam and the small diffraction angles due to the high forward velocity of the molecules. This method was also used for the analysis of physical properties of iodine I2 [49] and the potassium dimer K2 [88]. Picture reprinted from [228] with kind permission from Springer Science + Business Media.

Standard image High-resolution image

Because of the finite lifetime of the excited state only the black trajectory—with the molecular ground state in both arms—contributes to the final interference signal. The same short lifetime is also the reason why the spatial splitting between the two interferometer arms remains smaller than the width of the molecular beam and why it took the community many years to realize that this advanced spectroscopy scheme can also be seen as a genuine molecular matter-wave interferometer. Its key is nowadays used in metrology: applied to atoms with long-lived excited states this scheme implements the prototype of an optical atom clock [226]. Recent studies with potassium dimers 39K2 were successfully employed to demonstrate the measurement of a molecular transition dipole moment [227].

6.2. Mach–Zehnder experiments

One of the first closed interferometer for atoms (dashed line in figure 14) [229] was also used on molecular sodium (solid line). It served for the first demonstration of molecular interference with macroscopically split trajectories (several micrometres) and allowed the MIT team to measure the molecular polarizability both in an external electric field as well as in collision experiments [51]. The thermal particle beam was seeded into a supersonic source, sent through three nanomechanical gratings in Mach–Zehnder (MZI) geometry and detected by a hot-wire detector.

Figure 14.

Figure 14. MZI for Na2 [51]. One of the first interferometers for sodium atoms (dashed line) [229] was also used on sodium molecules (solid line) [51]. The two lasers shown in this scheme can be used to induce absorption, deflection and decoherence. Copyright 2013 by the American Physical Society.

Standard image High-resolution image

In an MZI configuration the gratings are positioned in the molecular far-field where the beam paths become physically separated. It is important that the initial collimation angle of the particle beam is smaller than the first-order diffraction angle at the nanogratings. The accessibility of the individual interferometer arms to lasers, static electric fields or even separate gas cells was used for metrology on atoms and molecules [51].

6.3. Near-field (Talbot–Lau) interferometry

6.3.1. The concept.

Mach–Zehnder interferometry with atoms and molecules was pioneering in matter-wave physics but also showed the challenges for using rare materials such as macromolecules. In the absence of coherent sources, far-field experiments require the preparation of spatial coherence, and thus in practice tight collimation. This is acceptable for intense atomic beams but it represents an almost insurmountable obstacle for macromolecular quantum optics—until new cooling schemes are developed. In the meantime, molecular quantum delocalization experiments can also be pursued in a near-field matter-wave interferometer.

The Talbot–Lau concept was already known from optics [204, 230] and it has been realized for the first time with atoms by Clauser [61] who also suggested that this idea would be a key to super-massive interferometry [62].

The implementation shown in figure 15 refers to a series of experiments with macromolecules performed in Vienna [63]. Molecules may arrive at the first transmission mask G1 without any initial spatial coherence. The comb of narrow slits pre-selects the possible starting points for the subsequent matter waves to evolve. All slits act individually as coherence-preparing elements. The tight localization in each slit of G1 imposes sufficient quantum uncertainty in the momentum of the particles to establish the transverse delocalization and coherence that is required to illuminate more than two neighbouring slits in the second grating. Quantum near-field interference then generates a particle density pattern at the position of G3, which is a self-image of G2 [62, 64, 231], as also seen in figure 8. If G3 is scanned across the interference pattern, the totally transmitted molecular flux will give a sine-like signal.

Figure 15.

Figure 15. The concept of molecular matter-wave interferometry in a TLI near-field configuration [63, 64]. The KDTLI and OTIMA configurations are derived from this concept. The TLI is transformed into a KDTLI, when G2 is replaced by an optical phase grating, i.e. a non-resonant standing light wave interacting with the molecular optical polarizability. The TLI can be converted into an OTIMA by replacing every nanomechanical mask G1–G3 by a pulsed and absorptive optical standing light wave, as for instance implemented by three vacuum ultraviolet laser pulses of several nanosecond duration and with equal time intervals between subsequent pulses. Copyright 2002 by the American Physical Society.

Standard image High-resolution image

The idea of TLI is nowadays used for verifying the wave–particle duality of massive molecules, for quantum-enhanced lithography and quantum-enhanced molecule metrology.

The key advantage of TLI over MZI is its high multiplexing capability. Without the need of collimation one can use broader molecular beams. The signal enhancement of three to five orders of magnitude was the enabling step for a multitude of novel experiments. At the same time, the scaling behaviour of near-field optics is favourable for the very short de Broglie wavelengths of very massive particles.

The Talbot–Lau setup is based on the idea that coherent self-imaging may occur when the diffraction orders of neighbouring slits overlap at the same positions on the screen. While the diffraction pattern originating from each individual starting point in G1 is a pure wave phenomenon, the mutual overlap between neighbouring interference patterns is a consequence of the experimental geometry.

This resonance condition is fulfilled if the gratings are positioned at the Talbot distance LT = d2/λdB, with d the grating period and λdB the de Broglie wavelength. It is also fulfilled at multiples of the Talbot length and it even leads to (weaker) fringes with a smaller spatial period at rational fractions of LT [204, 232].

The Talbot criterion implies that for a given experimental dimension—for instance limited by the lab size—the grating period only shrinks as $d\propto \sqrt \lambda_{{\rm dB}}$ . Two molecular species that travel at the same speed and differ in mass by a factor of 100 require only a ratio of 10 in their mask periods. For high-mass interferometry this is a key advantage compared with far-field interferometry where $d\propto \sqrt \lambda_{{\rm dB}}$ .

Over the years, different requirements led to a number of variations of this basic layout: first, it turns out that massive things are usually also highly polarizable. The dephasing van der Waals force that influenced earlier molecular far-field experiments is already detrimental for near-field interferometry with molecules in the range of 10 000 amu and grating openings of several dozen nanometres [65]. This is the reason why it proved necessary and successful to replace the central grating G2 in figure 15 by an optical phase grating. While the coherence preparation and fringe selection still have to be performed by absorptive gratings, here realized by the material masks G1 and G3, the diffraction is then caused by the Kapitza–Dirac effect, i.e. the interaction of delocalized matter with a non-resonant standing light wave. Originally proposed by Kapitza and Dirac for electrons [233] the effect of phase gratings was first seen for atoms [234], then with electrons [235] and molecules [60] in far-field diffraction. In combination with the near-field idea of the TLI it becomes the Kapitza–Dirac–Talbot–Lau interferometer (KDTLI) [65].

While the KDTLI interferometer still integrates the position-selective nanomechanical gratings G1 and G3 one can replace them also by absorptive structures of light. They can, for instance, be achieved when the photon energy exceeds the molecular ionization energy [66]. An interferometer based on three (typically ultraviolet) ionizing laser gratings can be operated both in space and in the time domain [67, 87]. An optical time-domain ionizing (OTIMA) interferometer is universal in the sense that it can be applied to a large variety of different particles, from single atoms to clusters of molecules. The quantum state in such interferometers may in the future also be retrieved by using state tomography [234].

6.3.2. Interferometric deposition of surface-bound molecular nanostructures.

Molecular TLI as shown in figure 15 intrinsically generates a nanosized molecular density pattern at a well-defined position behind the second grating. While in many experiments this pattern is sampled and visualized by scanning the third grating G3 across the molecular beam, it is also possible to capture the interference fringes on a screen to generate a surface-bound molecular nanostructure [42]. This idea has been pursued in a recent matter-wave experiment with C60 molecules, where STM imaging was used to visualize the resulting interference pattern after its deposition on a reconstructed Si (1 1 1) 7 × 7 surface. Such images also permit us to visualize the quantum wave–particle duality. In figure 16, every single spot represents a detected C60 molecule and although there is no way to predict a priori the position that any of them will have, the vertical ensemble average (bottom) shows the expected highly regular fringe pattern. And even though one would also expect a classical fringe shadow behind a two-grating arrangement [232], the expected and observed quantum fringe contrast are significantly higher—even in the presence of van der Waals forces. Quantum-enhanced molecule lithography may become relevant in nanotechnology as a positive and soft non-contact deposition technique for creating surface modifications on the nanoscale and with deposition energies as small as 0.1 eV.

Figure 16.

Figure 16. Scan of the surface-deposited molecular interference pattern behind a Talbot–Lau configuration with two mechanical SiN gratings and one Si-detection screen. Every single C60 molecule is depicted as a white dot. A vertical sum over the entire image yields the blue circles in the bottom part. The interference pattern is expected to be described by a sine curve, which was also used to fit the data [83]. Reprinted with permission from [42]. Copyright 2009 by the American Physical Society.

Standard image High-resolution image

6.3.3. Comoving fluorescence detector for near-field experiments.

Spatially highly resolving imaging methods, such as scanning tunnelling or single-molecule fluorescence microscopy require the suppression of molecular surface diffusion to less than a few ten nanometres. This is not granted for all molecule–surface combinations. But even then, position-encoded fluorescence detection can still be implemented using a mechanical scheme that magnifies the quantum fringe pattern to the extent that even sizeable surface diffusion can be completely neglected [237].

In a three-grating interferometer of the Talbot–Lau type (figure 17), the third grating, G3, can be regarded as part of the detector. It is scanned in the direction of the grating vector $\vec{k}=\vec{e}_{x}\cdot {2\pi}/d$ , i.e. normal to the molecular beam and normal to the extension of the diffraction slits. For each position, the transmitted molecular flux is then recorded.

Figure 17.

Figure 17. Quantum interference of tetraphenylporphyrin (TPP) detected in a comoving and mechanically magnifying fluorescence detection scheme. A local vertical average over the 2D image (red rectangles) allows one to retrieve the quantum fringes (right-hand graph, black dotted lines). We expect and find a sinusoidal shape of the fringe whose contrast varies with the particles' height on the screen, i.e. their most probable velocity. A = 33 µm (v ≈ 300 m s−1), B = 371 µm (v ≈ 270 m s−1), C = 938 µm (v ≈ 160 m s−1) and D = 1234 µm (v ≈ 140 m s−1). Reprinted from [237], with permission from IOP Publishing.

Standard image High-resolution image

In a recent experiment this method was combined with fluorescence imaging [237] of molecules that were accumulated on a quartz substrate mounted onto a translation stage behind G3. For each position shift of the third grating Δx3, the quartz plate was shifted by ∼4300Δx3. This trick magnifies the molecular interference pattern mechanically to an extent that it can be easily analysed and that molecular diffusion on the detection plate can be completely neglected. When using gravitational velocity selection, a 2D image, such as shown in figure 17 allows visualizing the velocity-dependent near-field matter-wave interferogram.

6.4. TLI for molecule metrology

All variants of TLI can be applied as tools for molecule metrology, i.e. the measurement of internal molecular properties [1]. The basic concept is to use the accurate spatial modulation of the molecular beam to observe the coupling of external fields on internal properties, which will cause a force that shifts the interference fringe pattern. The range of properties to be investigated is wide. Some experiments have already been performed to measure the static and dynamic polarizability [71] as well as the permanent or thermally activated electric dipole moment of molecules in the range of 1000 amu [39]. For that the shift of the interference pattern was monitored in the presence of an inhomogeneous electrostatic field (E∇) E. The deflection can, in principle, be used to spatially separate and sort pure molecular samples starting from a mixtures of species [38]. The conformational state dynamics of diazobenzene molecules could be mapped to the centre-of-mass motion by comparing the observed deflection with ab initio simulations [15]. Quantum deflection may thus provide benchmark data for increasingly sophisticated molecular simulations, both with regard to static properties or for instance optical absorption cross sections [40].

7. Prospects of molecular and nanoparticle interferometry

The present review focused mainly on experimental techniques that have been used with small molecules and methods that are required to establish quantum phenomena with large molecules and nanoparticles. In this outlook, we focus on the interest and possibilities in extending this approach to high mass and complexity.

7.1. Decoherence, phase averaging and fundamental limitations to nanoparticle interferometry

Molecule and nanoparticle interferometry is currently believed to be a valid and actually even one of the best possible approaches for testing the quantum superposition principle in a domain where a number of still speculative models have emerged over recent years with predictions of non-linear additions or other modification of quantum physics [35, 238].

In the following, we briefly discuss a number of practical and possibly fundamental limitations for quantum coherence experiments with massive bodies. We distinguish phase averaging mechanisms, genuine quantum decoherence and the possibility of subtle deviations from quantum linearity.

7.1.1. Decoherence.

A key distinguishing element in this comparison is the role of quantum information in the transition between observable quantum coherences and classical probabilities. Decoherence theory [21, 211, 212, 239242] is applied quantum mechanics and it describes the non-observability of coherence in a system as a result of the coupling to its complex physical environment: The exchange of quantum information, the diffusion of coherence into the environment by virtue of entangling interactions or 'measurements'—even without the need for invoking any human observer—is fully described by established quantum theory. It reduces quantum coherence effectively if we neglect to include all partners of the extended many-body system.

Understanding quantum decoherence is also central to the overall enterprise that is currently being pursued under the heading of quantum information technologies. Only if we know and master the effects that destroy quantum phenomena we can also take measures to extend coherence to an extent that enables us to build quantum-based sensors or information processing devices on the mesoscopic scale.

A number of matter-wave experiments were able to corroborate the ideas of decoherence theory: when particles are excited, such that each of them spontaneously emits a photon while they are delocalized inside a MZI (see figure 14), the overall quantum fringe contrast is reduced, if the emitted photon wavelength suffices to yield which-path information, i.e. information about where the particle was [243].

The 'Copenhagen complementarity' interpretation, which states that path information and quantum interference exclude each other, can nowadays be rephrased in terms of quantum information theory: information can only be extracted by a physical particle. If this scattering partner becomes quantum entangled with the originally delocalized nano-object, quantum coherence is diluted into the more complex environment.

In pure de Broglie interference, i.e. experiments which only consider the centre-of-mass motion, this information theoretical view can finally be complemented by a third perspective, namely the analogy to 'Heisenberg's microscope': when there is no coupling to the internal states, the only way to extract position information is momentum exchange. In that case, random kicks by the interacting environment can dephase each individual delocalized molecule differently to a degree that the coherence of the ensemble is finally lost. This effect depends on the separation of the wave packets as well as on the momentum exchange with the probing radiation [244].

In TLI with C60 decoherence was observed for the first time both caused by collisions with residual gases in the vacuum chamber [70, 245] and by thermal radiation emitted by the molecules themselves [22, 246]. Decoherence is also believed to be an important reason why molecules appear with a certain chirality in nature, instead of the energetically favoured superposition of left- and right-handed configurations [247].

The only way to circumvent decoherence is to isolate a particle from all external agents. In practice, this imposes rather stringent requirements on the background vacuum or temperature [35, 7476]. Extrapolations of particle interferometry with masses around 1010 amu are conceptually still compatible with existing technologies, demanding as they may be in practice [35, 74]. This will require cooling of the entire experiment to the temperature of liquid helium in order to reduce the base pressure and the level of black body radiation to a level acceptable for high-mass interference [248].

7.1.2. Phase averaging.

In contrast to decoherence, phase averaging may appear trivial: any advanced quantum experiment may suffer from many external perturbations, which do not invoke quantum information arguments and yet they impose very relevant boundary conditions to high-mass interference experiments. Some of the phase averaging mechanisms may even be related to fundamental physics from outside of quantum mechanics, such as gravitational wave background noise [249] or space–time fluctuations [250].

However, most causes for phase averaging are related to a much more basic and lab-oriented level. This includes dispersive effects in interferometry with molecular beams of finite velocity spread: the Coriolis-force on the Earth, the particle's free fall in the Earth's gravitational field, the van der Waals interaction with solid surfaces or the deflection in an inhomogeneous electric or magnetic field may all influence the phase of the quantum fringe very much in the same way as they shift a classical particle beam. Also different modes of interferometer vibrations will reduce the quantum contrast, without extracting quantum information [109].

The OTIMA concept [67, 68] eliminates a large number of dispersive effects and is therefore particularly promising for this kind of macro-interference. In addition, it starts from spatially and spectrally incoherent sources, which is a tremendous benefit, given the difficulties in preparing large ensembles of ultracold and similar nanoparticles. Also single-grating experiments become feasible again, once ultracold nanoparticle sources become available.

7.1.3. Is quantum mechanics all there is?

In neither one of the two above arguments the superposition principle is ever broken. Conceptually, this leads either to an ever-growing entanglement in the universe or a many-world interpretation of quantum mechanics [251].

This is the reason why a number of proposals have been made in the attempt to establish 'reality', i.e. physical states without the superposition of classically mutually exclusive phenomena. This has for instance been done by adding non-linearities to the Schrödinger equation [17, 19, 252] or by exploring the relation of quantum mechanics in combination with gravity theory [18, 36, 37].

This summarizes to a picture where quantum mechanics, at present, is the best-confirmed theory of non-relativistic physics—and even better tested in its relativistic form of quantum electrodynamics—but where a growing number of models emerge to explore potential limits and modifications of the superposition principle for very massive bodies.

8. Conclusion

This report has been written to sketch the present state of the art in matter-wave interferometry with particles more complex than single atoms. We have discussed several experimental techniques to prepare, manipulate and detect molecular beams. Applications in mass spectroscopy, in surface studies, explorations of van der Waals forces, characterization of molecular bonds and energies as well as demonstrations of fundamental quantum phenomena with complex particles show the wide applicability of mesoscopic matter waves.

Coherent optics with large molecules is still a young field and we expect major progress in future years: more sophisticated cooling techniques and coherent beam sources will open new possibilities in manipulating complex nanoparticles. Coupling internal with external degrees of freedom is expected to lead to new forms of spectroscopy. Improved coherent control of motional states shall take us to coherent molecular microscopy and lithography. Modern experiments in molecule interferometry cover de Broglie wavelengths between 100 fm and 100 µm with energies ranging between 100 neV and 10 keV.

Advances in the manipulation of single molecules open a new reductionist way to molecular beam physics with full quantum control at the level of individual molecular systems.

The advent of measurement-induced gratings made of light has led to new experiments that will allow testing quantum theory in an unprecedented mass range. Several breakthroughs throughout recent years have given credibility to nanoparticle interferometry and the possibility to quantitatively test the linearity of quantum physics in a new domain.

A fascinating aspect of quantum physics with complex compounds is the seemingly unlimited number of test objects, ranging from tailored inorganic materials, over metals and semiconductor clusters or nanospheres up to biomolecules and nanomaterials at the interface to life. This opens the access to an unprecedented wealth of interaction possibilities to tailor and tune the particles to the needs of the experiments.

Quantum-interference-assisted metrology, based on the coupling of internal and external degrees of freedom can be exploited for new measurements of electronic, magnetic, optical or structural properties, and the exchange of quantum information between internal and external states may open a new avenue to fundamental quantum studies.

We conclude by expressing the desire that many partners will join us in our research. The complexity of the subject is enormous and many hands and brains will be needed to harvest the fruits that grow on the wide field of molecular and nanoparticle quantum optics.

Acknowledgments

MA owes special thanks to Anton Zeilinger with whom we started the first quantum experiments with macromolecules and who has remained a driving force in questioning the status quo of knowledge. We thank our co-workers at the University of Vienna and the University of Southampton as well as our collaboration partners, in particular the groups around Klaus Hornberger, Marcel Mayor, Ori Cheshnovsky, Uzi Even, Angelo Bassi and Markus Aspelmeyer, for their work and inspiration that make matter-wave optics with clusters, molecules and nanoparticles an exciting adventure. We thank the Austrian Science Fund, FWF for financial support in the projects Z149-N16 (Wittgenstein) as well as the European Research Council in project (ERC AdG 320694 PROBIOTIQUS) and the European Commission in project (304886 NANOQUESTFIT). HU thanks the UK funding agency EPSRC (EP/J014664/1), the Foundational Questions Institute (FQXi-RFP3-1021) and the John F Templeton foundation (ID 39530) for support. TJ thanks the Gordon and Betty Moore Foundation for support.

Please wait… references are loading.