Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

Refined model for Talbot–Lau matter-wave optics with pulsed photodepletion gratings

Open Access Open Access

Abstract

We analyze time-domain Talbot–Lau interferometry of organic cluster beams that are exposed to pulsed photodepletion gratings in the vacuum ultraviolet. We focus particularly on the analysis of the complex (phase and absorption) character of the optical elements. The discussion includes the role of wavefront distortions due to mirror imperfections on the nanometer level and the effect of finite coherence in the diffraction gratings. This improved understanding of the interferometer allows us to extract new information on optical properties of anthracene and ferrocene clusters and to define conditions for future matter-wave experiments.

© 2014 Optical Society of America

1. INTRODUCTION

Talbot–Lau interferometry has attracted significant attention in the optics community, since it allows one to coherently generate self-images of periodic structures using spatially incoherent light sources. This is of practical relevance in classical light optics [1,2], and it opens new avenues in fields of research in which coherent sources or refractive optical elements are not readily available. This holds for x-ray physics [3,4], plasmonics [5], and ultrasound imaging [6]. For the same reason Talbot–Lau interferometry was proposed to be a promising tool for matter waves, particularly for massive objects [7]. The idea has been successfully demonstrated in experiments with electrons [8], atoms [912], and hot molecules [13,14].

Many of these earlier Talbot–Lau experiments were built using microfabricated diffraction gratings. However, complex particles of high polarizability are best manipulated using masks of light. This avoids the perturbing influence of van der Waals interactions or static charges that occur when particles transit material masks with slits as narrow as 100 nm [15,16,13]. In response to the need, standing lightwave phase gratings [17,18] and absorptive photoionization gratings [19] as well as photofragmentation gratings [20] have been introduced to quantum optics experiments with macromolecules and molecular clusters. Especially, pulsed optical diffraction elements offer high versatility since the light–matter interaction can be timed with nanosecond precision and tuned via the laser wavelength, energy, or pulse length.

Here, we revisit the concept of optical time-domain matter-wave interferometry (OTIMA), which uses the periodic selection, diffraction, and detection of nanoparticles in a series of three pulsed standing lightwave gratings [19,21]. We focus on the practical limits imposed by the specifications of available laser systems and optical components at a wavelength of 157 nm. Particular attention is paid to the influence of a running wave that is superimposed to each standing lightwave grating. This offset is caused by the finite laser coherence and the mirror reflectivity. We further study the relevance of the local mirror roughness and demonstrate how to cope with global mirror surface deformations (flatness) on the level of 10 nm. The improved characterization of the OTIMA interferometer allows us to gain new information about optical cross sections and polarizabilities of particles such as molecular clusters of anthracene and ferrocene.

2. EXPERIMENTAL LAYOUT

The experimental sequence is as follows: molecules are evaporated inside an Even–Lavie valve [22] in the presence of high-pressure seed gas (1–7 bar of neon). A nozzle releases short (30 μs) dense pulses of molecules that subsequently cool and cluster in the adiabatic expansion. With a mean velocity of around 900ms1, the particle beam is skimmed, collimated, and transmitted into the interferometer chamber via two pumping stages. In our present work, we focus on clusters of anthracene (AcN=(C14H10)N) and ferrocene (FcN=(C10H10Fe)N), since their volatilization and ionization properties are well suited for our analysis. For the measurements presented here, the molecular samples were heated to 430 and 500 K, respectively. A typical mass distribution of ferrocene clusters is shown in Fig. 1(b).

 figure: Fig. 1.

Fig. 1. Setup for cluster interferometry using three pulsed, absorptive standing lightwave gratings. (a) Mirror deformations shift the nodes of the standing waves within the laser spot. This reduces the fringe visibility due to averaging over the phase-shifted interference patterns. This is indicated by the solid and dashed semiclassical paths for two particles that start with the same velocity and direction but at different positions. A mirror reflectivity R<0.96 and limited laser coherence are the reasons why a running wave overlays the periodic gratings. The figure is drawn not to scale to illustrate the effects of minuscule mirror deformations. (b) Mass spectrum of ferrocene clusters that are detected after the third grating pulse using VUV photoionization in a time-of-flight mass spectrometer. We compare the interference signal SInt (blue dashed line) with a reference signal SRef (red solid line), as shown in the inset for the model system Fc7. (c) The normalized signal contrast SN (see text) is shown for the clusters Fc3 to Fc9.

Download Full Size | PDF

OTIMA interferometry, as depicted in Fig. 1(a), combines three VUV photodepletion gratings (G1G3). These standing lightwave gratings are formed by retroreflecting three fluorine (F2) excimer laser beams (λ=157.6nm, E3mJ, τ=8ns) from the surface of a single 5 cm diameter CaF2 dielectric mirror. Each beam has a flat-top profile focused to an area of A=10mm×1mm, and neighboring beams are separated by 20 mm. The direction of the elongated waist is aligned with the cluster beam. At a photon energy of 7.9 eV (157 nm), the clusters efficiently ionize or fragment after absorption of a single photon. Since both processes deplete the particle beam, the gratings represent absorptive (beam depletion) masks with a period of d=λ/2=78.8nm. Clusters that survive the interferometer sequence are post-ionized with a fourth F2 laser pulse and detected by time-of-flight mass spectrometry. This allows us to compare particles with different masses and optical properties in one and the same experimental run.

The absorptive character of grating G1 confines the molecules in a comb of starting positions, so that their wave functions expand over several nodes and antinodes of the standing lightwave by the time the second grating is triggered. Rephasing of the wavelets behind G2 leads to multipath interference and to the formation of a cluster density pattern, which can be resolved with the third grating pulse G3. This final pulse samples the interference pattern and transmits only those clusters whose wave function is aligned with the grating nodes. As in other realizations of Talbot–Lau interferometry, the coherent self-imaging of the diffraction grating occurs on a characteristic time scale, the Talbot time TT=md2/h, where m is the particle’s mass, d the grating period, and h Planck’s constant.

We measure the fringe visibility by recording the transmitted number of particles behind G3 as a function of the cluster mass. For this, we choose a fixed pulse separation time T between two grating pulses and use two complementary measurement settings to visualize interference. In the interference setting, the cluster signal SInt is recorded, while the two pulse separation times T=T3T2=T2T1 are kept equal to within ΔT<1ns. Interference then enhances or reduces (depending on the phase) the cluster signal for masses close to the Talbot criterion. In the reference mode (SRef), we set an imbalance in the pulse separation times by ΔT>50ns. For the given divergence and mass range in the molecular beam, this is sufficient to wash out the interference pattern [23].

We define the cluster interference contrast as the normalized signal difference SN=(SIntSRef)/SRef, with SN [1,1] [Fig. 1(c)]. In the hypothetical case of a perfectly flat mirror surface and in the absence of external accelerations, SN is positive and equal to the theoretical visibility V of the interference pattern (see Section 4). For a deformed mirror [see Fig. 1(a)], however, SN is given by the visibility of the cluster density pattern at G3 and the relative displacement ΔD=Δx12Δx2+Δx3 of this Talbot image with respect to G3. Maximal constructive interference occurs for ΔD=zd and maximal destructive interference for ΔD=(z+1/2)d, with zZ. The grating shifts Δxi (i=1,2,3) refer to a common zero line. In the present experiments, the corrugated mirror surface leads to negative SN. Inertial or electromagnetic accelerations a may introduce additional fringe displacements, which are, however, negligible for the parameters of our present experiments. If the mirror surface deformation exceeds 5–10 nm in the small effective region where the light pulse and the cluster cloud interact, different particles will experience differently shifted interferometers, as indicated by the two semiclassical paths in Fig. 1(a). The effective interaction region is set by the width of the detecting laser beam, i.e., 1–2 mm. Averaging will, thus, lead to reduced visibility of the density pattern at the time of G3. We account for the shift and reduction of the fringes by multiplying the theoretical visibility V [Eq. (7)] by the fringe factor Ω [1,1], so that SN=Ω·V. On the one hand, this factor contains the interferogram shift for a single particle relative to G3. On the other hand, Ω takes into account the reduction of the visibility due to averaging overshifted interferograms. It is determined by the mirror geometry in combination with the chosen pulse timings T1, T2, T3 and the length of the detected cluster cloud.

3. MODELING FINITE MIRROR REFLECTIVITY AND FINITE LASER COHERENCE

The 1D intensity profile of a standing lightwave along x is ideally described by I(x)=4I0cos2(2πx/λ), with I0=E/(Aτ) the incident laser intensity. In reality, however, we need to take into account imperfections of the interferometer mirror and the grating lasers. At VUV wavelengths around 157 nm, state-of-the-art CaF2 mirrors are limited to a reflectivity of R=96%. This may further degrade in the presence of hydrocarbon contaminations [24]. In addition, our free-running F2 lasers (EX 50, Gam Laser Inc.) have a limited longitudinal and transverse coherence. They are expected to emit dominantly at 157.63 nm (ca. 90% of the total intensity) and 157.52 nm (ca. 10%) with a linewidth of Δλ1pm [25]. This corresponds to a longitudinal coherence length of about 1 cm. The transverse coherence is determined by the width of the emitting aperture and amounts to 50–100 μm at the location of our interferometer mirror in 2 m distance from the laser exit.

These imperfections add an incoherent sum of running waves to the standing wave light field, which reduces visibility Vλ of the laser grating. This reduces, in turn, the expected interference contrast of the diffracted matter-wave in dependence of the particle’s VUV absorption cross section. Inhomogeneities within each grating, due to local roughness of the mirror surface, cause each individual particle to average over different standing wave phases during the laser pulse. Similarly, a divergent or tilted (with respect to the mirror surface) particle beam as well as mirror vibrations can lead to averaging over grating phases during the 8 ns pulse. For example, particles with a velocity of 1000ms1 orthogonal to the grating vector will accumulate over 8 μm of the standing light field. Thus, a particle beam with a divergence of 1 mrad will average the field over 8 nm in the grating direction. These effects strongly depend on the absorption cross section of each particle at the laser wavelength λ.

To account for this running wave contribution in our model, we distribute the incident laser intensity among a one-directional nonreflected contribution, I(x)=(1R)I0, two counterpropagating running waves I(x)=2R(1C)I0, and the standing lightwave 4RCI0cos2(2πx/λ). We quantify the coherent contribution by C [0,1]:

I(x)=(1R)I0+R[2(1C)I0+4CI0cos2(2πxλ)].

The transmission function t is given by the cluster survival probability, which is the probability of not absorbing any photon during the grating pulse. The mean number of absorbed photons per pulse depends on the optical absorption cross section σλ at λ=157nm. It is given by n(x)=I(x)λσλτ/(hc). Assuming Poissonian statistics and averaging the transmission function t over one grating period, we obtain

t=2λ0λ/2exp[n(x)]dx=J0(in02)exp[n0(14RC+14C)]J0(in02)exp(n02Vλ).
Here, J0 denotes the zeroth-order Bessel function, and we introduce n0=4RCI0λσλτ/(hc) as the difference between the mean number of photons absorbed in the antinodes and the nodes of the standing wave. The visibility Vλ of the light grating is given by Vλ=2RC/(1+R). If Vλ is known, we can extract n0(i) for every individual grating and all cluster masses separately by measuring the cluster transmission t(i).

4. INFLUENCE OF THE SECOND GRATING

The established interferometer theory [26] describes the fringe visibility as a function of the molecular beam depletion in all three gratings and includes the dipole interaction in G2 between the particle’s optical polarizability αλ and the electric laser light field. This allows us, in principle, to extract the particle’s optical properties from the interference contrast as a function of the laser power in G2.

In a purely absorptive grating, the difference between the mean number of absorbed photons in the antinodes and the nodes of the standing wave n0 determines the effective open fraction of the grating and the interference contrast. The action of the first and third grating in a Talbot–Lau interferometer is fully characterized by n0(1) and n0(3).

In the second grating, the electric field additionally imprints a relevant spatially dependent phase shift onto the traversing matter waves, which is given by

Φ(x)=4π2αλRCI0τhccos2(2πxd)Φ0(2)cos2(2πxd),
where Φ0(2)=4π2αλRCI0τ/(hc) is the relative phase shift between two parts of the wave that pass the antinode and the node of G2, respectively.

It is convenient to introduce the dimensionless parameter β as the ratio of the absorption cross section and the optical polarizability αλ:

β=λσλ8π2αλ=n0(i)2Φ0(i).

The signal behind the third grating in a symmetric Talbot–Lau configuration, as shown in Fig. 1, can be Fourier expanded and expressed as a function of the relative grating shift ΔD (see [21] for details):

S(ΔD)=l=Bl(1)(0)B2l(2)(lTTT)Bl(3)(0)×exp[2πild(ΔDaT2)],
where a denotes external acceleration along the grating direction (negligible in the present experiments), and Bn(i) are the Fourier coefficients that describe the influence of the individual gratings and are given by
Bn(i)(ξ)=exp(n0(i)/2)(sinπξβcosπξsinπξ+βcosπξ)n/2×Jn[sgn(1βsinπξ+cosπξ)×n0(i)21β2sin2πξcos2πξ],
with the Bessel functions Jn.

The visibility of the fringe pattern is then defined as

V=maxx[S(x)]maxx[S(x)]maxx[S(x)]+maxx[S(x)][0,1].

If the pulse separation time fulfills the Talbot criterion T=l·TT (with lN), the interference pattern is unaffected by the phase contribution, since all terms containing β in Eq. (6) drop out. In this case, the visibility V is only determined by n0(i) of the three gratings, as seen in Fig. 2. For symmetric pulse delays deviating from the Talbot time, T2T1=T3T2l·TT, the optical polarizability αλ influences the interference contrast. The time dependence of the contrast would theoretically allow us to extract β from such curves. Experimentally, however, changes of the time delay cause changes of the grating phases because of the mirror curvature and corrugation. In practice, measuring the contrast as a function of the laser energy at a fixed delay is more robust and allows us to retrieve β from a fit to the data.

 figure: Fig. 2.

Fig. 2. Theoretical visibility V for different values of the optical polarizability. (a) Contrast as a function of the pulse separation time T, in multiples of the Talbot time TT. It is plotted for varying phase shifts (polarizabilities) in the antinodes of the second standing wave (Φ0(2)=7.5, dotted; Φ0(2)=3, dashed; Φ0(2)=1.5, solid). (b) Contrast as a function of Φ0(2), i.e., for varying optical polarizability αλ, plotted for three different pulse separation times (T=0.75TT, dashed; T=0.9TT, dotted, T=TT, solid). For T=TT, the particle polarizability does not affect the interference contrast at all, while at 0.75TT the contrast oscillates with Φ0(2). The arrows and colors link plot (a) with plot (b). For both panels, the transmission through all three gratings was fixed by setting n0(i)=6.

Download Full Size | PDF

5. EXPERIMENTAL RESULTS

In order to extract the optical properties for all clusters, several parameters need to be measured in the experiment: An overall reduction of the interference contrast due to mirror deformations, common to all cluster numbers N, is included in the factor Ω. The visibility Vλ is assumed to be the same for all gratings, since the laser beams are equally well aligned to the same mirror surface, and the laser properties are known to be insensitive to the detailed laser conditions, such as its gas pressure or temperature [25]. The cluster interference contrast will, however, depend on Vλ since the grating quality influences the open fraction of the gratings n0(i). Since the cluster size and structure are relevant, the parameter β must be fitted for every cluster number, individually.

In order to extract these parameters from our measurements, we proceed as follows: We record the normalized contrast SN for all clusters in the accessible mass spectrum simultaneously as a function of the laser power in G2. To attenuate the laser energy, we use the VUV absorption of oxygen [27] in a pressure cell filled with air. This cell is inserted into the evacuated beam guiding system, separated by two CaF2 windows. It allows adjusting the transmitted laser power from a few percent up to 90% of the power at the cell entrance window, without compromising the timing performance or the beam profile. The laser pulse separation time T is set to the Talbot time of the hexamer (T=17.38μs for ferrocene Fc6, T=16.65μs for anthracene Ac6) such that high contrast is expected around these masses (see Fig. 1). In addition, we record the transmission t(i) through every grating, individually, and feed this information into the theoretical model (Sections 3 and 4; see also [21]). The free parameters are adjusted until the theoretical predictions fit the experimental data. We start with one value of Ω and fit Vλ for the cluster that fulfills the Talbot criterion (the hexamer). This contrast depends only on the absorption in the standing wave (n0(i)), not on the optical polarizability αλ, as shown in Fig. 2. The resulting visibility Vλ and the measured grating transmissions are used to calculate n0(i) with Eq. (2) for all cluster numbers N. With these values of n0(i), the individual β is fitted for those clusters whose Talbot time does not match the chosen pulse separation time (here, Fc7, Fc8, Fc9 and Ac7, Ac8, Ac9). The minimization of the least-squares fit in Vλ and all β values is repeated for varying Ω to find the best match for all parameters. The results of this fitting procedure are shown in Fig. 3 for anthracene and Fig. 4 for ferrocene clusters. The routine converges very well, and we find good agreement between the model and the experimental data.

 figure: Fig. 3.

Fig. 3. Measured interference contrast as a function of the mean number n0(2) of standing wave photons absorbed in the antinode of the second grating. (a) Interference measurements for three different laser energies reveal the shift of the maximal interference contrast toward higher masses for lower laser pulse energy. The triangles link the plots (a) and (b). (b) Normalized contrast for anthracene clusters of size N=6–9: Ac6 (gray, dashed–dotted), Ac7 (red, dotted), Ac8 (light red, solid), Ac9 (brown, dashed). The lines are fits based on the model discussed in the text. Error bars represent one standard deviation of statistical error.

Download Full Size | PDF

 figure: Fig. 4.

Fig. 4. Quantum interference of ferrocene clusters from N=69. It is shown as a function of number of absorbed photons in G2, which is also a measure for the open fraction of the grating. Fc6 (gray, dashed–dotted fit), Fc7 (dark blue, dotted fit), Fc8 (light blue, solid fit), Fc9 (blue, dashed fit). Error bars represent one standard deviation of statistical error.

Download Full Size | PDF

Figure 3(a) displays the normalized contrast SN for clusters between N=39, where the laser pulse energy in G2 is varied to yield an average n0(2) in the octamer Ac8 of 3.5 (left panel), 2.3 (middle panel), and 1.6 (right panel). Even though the pulse delay time was set to match the Talbot time of Ac6, we find the highest contrast at higher masses. This agrees with the expected shift of the resonance as a function of the particle polarizability, and it includes the contribution of the running wave background. Figure 3(b) traces SN as a function of the absorbed photon number in G2, for clusters composed of N=6,7,8,9 anthracene molecules.

The results of the model fit are collected in Table 1. We find a value Vλ=0.75, indicating high quality of the standing lightwaves at a distance of 1 mm to the mirror surface. Repeating the same experimental sequence with identical source settings, and the mirror retracted to 2 mm to the cluster beam, we find a decrease in the light grating visibility to Vλ=0.70. This is consistent with a coherence length of the fluorine lasers of about 1 cm. The overall reduction of the fringe visibility by almost a factor of two compared with the ideal theory value shows there is still room for development of high-quality VUV optics. The reduced contrast is partially due to the finite mirror reflectivity (R<0.96). Even larger influence is attributed to the finite mirror curvature. Independent measurements on different mirror species revealed a global mirror curvature of about 6 km, bending the surface by up to 100 nm over 25 mm. Even this tiny displacement across the laser-cluster interaction zone (Fig. 1) may cause measurable phase averaging. To the first order, one might expect β to be independent of the cluster number N. For localized and noninteracting electrons, the absorption cross section and the polarizability should grow with the number of molecular constituents and their participating electrons. This assumption is essentially confirmed by Table 1. One may see a small decrease in β from N=7 to N=9, both in the case of anthracene and ferrocene. But the data are also compatible with the observation that the β parameters vary by more than the one sigma systematic error of the values in Table 1 when the experiment is repeated daily with changing source conditions. Depending on the source properties, such as its opening time, temperature, and the conditions of the adiabatic expansion, the average cluster conformation, absorption cross section, and polarizability may vary [28]. For similar source and interferometer settings, the β parameters of anthracene were, however, found to be robust on the level of 10%–20%, in all recent experiments. There is good agreement between the model and many dozens of data points, as well as the reproducibility of the β parameters.

Tables Icon

Table 1. Optical Parameters of Laser Grating and Molecular Clustersa

Based on this successful concept, we can also analyze the behavior of ferrocene clusters in OTIMA interferometry, as shown in Fig. 4. The data quality is comparable with the case of anthracene: the model fits the data well and allows us to determine the optical cluster parameters.

6. CONCLUSION

We have presented new analysis of matter-wave interferometry with pulsed photodepletion gratings in the vacuum ultraviolet regime. Our description takes into account real-world experimental limitations of the mirror flatness and roughness as well as the limited reflectivity of the VUV optics. This is essential for better understanding of the OTIMA concept, which is a generic idea for matter-wave interferometry with complex nanoparticles. Our method allows us to extract for the first time quantitatively the β parameter of organic clusters, i.e., the ratio of their absorption cross section and polarizability at 157 nm. This is remarkable, as it is generally demanding to assess optical properties of weakly bound van der Waals clusters in the vacuum ultraviolet.

Matter-wave experiments have been proposed to test the linearity of quantum mechanics with massive particles [26,2932]. Such studies require the best possible understanding of all mechanisms, which may cause an experimental deviation from theoretical predictions. This includes quantum decoherence in the presence of molecular collisions, scattering, or emission of photons [14]. The good quantitative description that could be achieved here underlines that it is promising to further proceed with explorations of high-mass quantum interference.

ACKNOWLEDGMENTS

We acknowledge support by the European Commission (NANOQUESTFIT, 304886), the European Research Council (PROBIOTIQUS, 320694), and the Austrian science funds (DK CoQuS W1210-3). We thank M. Schulz (PTB, Braunschweig), J. Ebert (Laser Optik Garbsen), and T. Klein (Bruker Nano) for characterizations of CaF2 optics.

REFERENCES

1. K. Patorski, The Self-Imaging Phenomenon and Its Applications (Elsevier, 1989), pp. 2–108.

2. W. A. Lohmann and D. E. Silva, “A Talbot interferometer with circular gratings,” Opt. Commun. 4, 326–328 (1972). [CrossRef]  

3. T. Weitkamp, A. Diaz, C. David, F. Pfeiffer, M. Stampanoni, P. Cloetens, and E. Ziegler, “X-ray phase imaging with a grating interferometer,” Opt. Express 13, 6296–6304 (2005). [CrossRef]  

4. F. Pfeiffer, T. Weitkamp, O. Bunk, and C. David, “Phase retrieval and differential phase-contrast imaging with low-brilliance x-ray sources,” Nat. Phys. 2, 258–261 (2006). [CrossRef]  

5. M. Dennis, N. Zheludev, and F. Garc de Abajo, “The plasmon Talbot effect,” Opt. Express 15, 9692–9700 (2007). [CrossRef]  

6. V. M. Ginzburg and A. M. Andreyev, “Small static and dynamic target detection in nontransparent medium and opaque water,” Proc. SPIE 3373, 11–15 (1998).

7. J. Clauser, De Broglie-Wave Interference of Small Rocks and Live Viruses (Kluwer Academic, 1997), pp. 1–11.

8. B. J. McMorran and A. D. Cronin, “An electron Talbot interferometer,” New J. Phys. 11, 033021 (2009). [CrossRef]  

9. J. F. Clauser and S. Li, “Talbot–von Lau atom interferometry with cold slow potassium,” Phys. Rev. A 49, R2213 (1994). [CrossRef]  

10. L. Deng, E. W. Hagley, J. Denschlag, J. Simsarian, M. Edwards, C. Clark, K. Helmerson, S. Rolston, and W. Phillips, “Temporal, matter-wave-dispersion Talbot effect,” Phys. Rev. Lett. 83, 5407–5411 (1999). [CrossRef]  

11. S. B. Cahn, A. Kumarakrishnan, U. Shim, T. Sleator, P. R. Berman, and B. Dubetsky, “Time-domain de Broglie wave interferometry,” Phys. Rev. Lett. 79, 784–787 (1997). [CrossRef]  

12. M. J. Mark, E. Haller, J. G. Danzl, K. Lauber, M. Gustavsson, and H.-C. Nägerl, “Demonstration of the temporal matter-wave Talbot effect for trapped matter waves,” New J. Phys. 13, 085008 (2011). [CrossRef]  

13. B. Brezger, L. Hackermüller, S. Uttenthaler, J. Petschinka, M. Arndt, and A. Zeilinger, “Matter-wave interferometer for large molecules,” Phys. Rev. Lett. 88, 100404 (2002). [CrossRef]  

14. K. Hornberger, S. Gerlich, P. Haslinger, S. Nimmrichter, and M. Arndt, “Colloquium: Quantum interference of clusters and molecules,” Rev. Mod. Phys. 84, 157–173 (2012). [CrossRef]  

15. R. E. Grisenti, W. Schöllkopf, J. P. Toennies, G. C. Hegerfeldt, and T. Köhler, “Determination of atom-surface van der Waals potentials from transmission-grating diffraction intensities,” Phys. Rev. Lett. 83, 1755 (1999). [CrossRef]  

16. V. P. A. Lonij, C. E. Klauss, W. F. Holmgren, and A. D. Cronin, “Atom diffraction reveals the impact of atomic core electrons on atom-surface potentials,” Phys. Rev. Lett. 105, 233202 (2010). [CrossRef]  

17. O. Nairz, B. Brezger, M. Arndt, and A. Zeilinger, “Diffraction of complex molecules by structures made of light,” Phys. Rev. Lett. 87, 160401 (2001). [CrossRef]  

18. S. Gerlich, L. Hackermüller, K. Hornberger, A. Stibor, H. Ulbricht, M. Gring, F. Goldfarb, T. Savas, M. Müri, M. Mayor, and M. Arndt, “A Kapitza–Dirac–Talbot–Lau interferometer for highly polarizable molecules,” Nat. Phys. 3, 711–715 (2007). [CrossRef]  

19. P. Haslinger, N. Dörre, P. Geyer, J. Rodewald, S. Nimmrichter, and M. Arndt, “A universal matter-wave interferometer with optical ionization gratings in the time domain,” Nat. Phys. 9, 144–148 (2013). [CrossRef]  

20. N. Dörre, J. Rodewald, P. Geyer, B. v. Issendorff, P. Haslinger, and M. Arndt, “Photofragmentation beam splitters for matter-wave interferometry,” arXiv:quant-ph/1407.0919 (2014).

21. S. Nimmrichter, P. Haslinger, K. Hornberger, and M. Arndt, “Concept of an ionizing time-domain matter-wave interferometer,” New J. Phys. 13, 075002 (2011). [CrossRef]  

22. U. Even, J. Jortner, D. Noy, N. Lavie, and C. Cossart-Magos, “Cooling of large molecules below 1 K and He clusters formation,” J. Chem. Phys. 112, 8068–8071 (2000). [CrossRef]  

23. M. Arndt, N. Dörre, S. Eibenberger, P. Haslinger, J. Rodewald, K. Hornberger, S. Nimmrichter, and M. Mayor, “Matter-wave interferometry with composite quantum objects,” in Atom Interferometry, G. Tino and M. Kasevich, eds. (IOS, 2014).

24. A. K. Bates, M. Rothschild, T. M. Bloomstein, T. H. Fedynyshyn, R. R. Kunz, V. Liberman, and M. Switkes, “Review of technology for 157-nm lithography,” IBM J. Res. Dev. 45, 605–614 (2001). [CrossRef]  

25. C. J. Sansonetti, J. Reader, and K. Vogler, “Precision measurement of wavelengths emitted by a molecular fluorine laser at 157 nm,” Appl. Opt. 40, 1974–1978 (2001). [CrossRef]  

26. S. Nimmrichter, Macroscopic matter-wave interferometry, thesis (Springer, 2014).

27. K. Yoshino, W. Parkinson, K. Ito, and T. Matsui, “Absolute absorption cross-section measurements of Schuhmann-Runge continuum of O2 at 90 and 295 K,” J. Mol. Spectrosc. 229, 238–243 (2005). [CrossRef]  

28. F. Piuzzi, I. Dimicoli, M. Mons, P. Millie, V. Brenner, Q. Zhao, B. Soep, and A. Tramer, “Spectroscopy, dynamics and structures of jet formed anthracene clusters,” Chem. Phys. 275, 123–147 (2002). [CrossRef]  

29. S. Nimmrichter and K. Hornberger, “Macroscopicity of mechanical quantum superposition states,” Phys. Rev. Lett. 110, 160403 (2013). [CrossRef]  

30. S. Nimmrichter, K. Hornberger, P. Haslinger, and M. Arndt, “Testing spontaneous localization theories with matter-wave interferometry,” Phys. Rev. A 83, 043621 (2011). [CrossRef]  

31. S. L. Adler and A. Bassi, “Is quantum theory exact?” Science 325, 275–276 (2009). [CrossRef]  

32. A. Bassi, K. Lochan, S. Satin, T. P. Singh, and H. Ulbricht, “Models of wave-function collapse, underlying theories, and experimental tests,” Rev. Mod. Phys. 85, 471–527 (2013). [CrossRef]  

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (4)

Fig. 1.
Fig. 1. Setup for cluster interferometry using three pulsed, absorptive standing lightwave gratings. (a) Mirror deformations shift the nodes of the standing waves within the laser spot. This reduces the fringe visibility due to averaging over the phase-shifted interference patterns. This is indicated by the solid and dashed semiclassical paths for two particles that start with the same velocity and direction but at different positions. A mirror reflectivity R < 0.96 and limited laser coherence are the reasons why a running wave overlays the periodic gratings. The figure is drawn not to scale to illustrate the effects of minuscule mirror deformations. (b) Mass spectrum of ferrocene clusters that are detected after the third grating pulse using VUV photoionization in a time-of-flight mass spectrometer. We compare the interference signal S Int (blue dashed line) with a reference signal S Ref (red solid line), as shown in the inset for the model system Fc 7 . (c) The normalized signal contrast S N (see text) is shown for the clusters Fc 3 to Fc 9 .
Fig. 2.
Fig. 2. Theoretical visibility V for different values of the optical polarizability. (a) Contrast as a function of the pulse separation time T , in multiples of the Talbot time T T . It is plotted for varying phase shifts (polarizabilities) in the antinodes of the second standing wave ( Φ 0 ( 2 ) = 7.5 , dotted; Φ 0 ( 2 ) = 3 , dashed; Φ 0 ( 2 ) = 1.5 , solid). (b) Contrast as a function of Φ 0 ( 2 ) , i.e., for varying optical polarizability α λ , plotted for three different pulse separation times ( T = 0.75 T T , dashed; T = 0.9 T T , dotted, T = T T , solid). For T = T T , the particle polarizability does not affect the interference contrast at all, while at 0.75 T T the contrast oscillates with Φ 0 ( 2 ) . The arrows and colors link plot (a) with plot (b). For both panels, the transmission through all three gratings was fixed by setting n 0 ( i ) = 6.
Fig. 3.
Fig. 3. Measured interference contrast as a function of the mean number n 0 ( 2 ) of standing wave photons absorbed in the antinode of the second grating. (a) Interference measurements for three different laser energies reveal the shift of the maximal interference contrast toward higher masses for lower laser pulse energy. The triangles link the plots (a) and (b). (b) Normalized contrast for anthracene clusters of size N = 6 –9: Ac 6 (gray, dashed–dotted), Ac 7 (red, dotted), Ac 8 (light red, solid), Ac 9 (brown, dashed). The lines are fits based on the model discussed in the text. Error bars represent one standard deviation of statistical error.
Fig. 4.
Fig. 4. Quantum interference of ferrocene clusters from N = 6 9 . It is shown as a function of number of absorbed photons in G 2 , which is also a measure for the open fraction of the grating. Fc 6 (gray, dashed–dotted fit), Fc 7 (dark blue, dotted fit), Fc 8 (light blue, solid fit), Fc 9 (blue, dashed fit). Error bars represent one standard deviation of statistical error.

Tables (1)

Tables Icon

Table 1. Optical Parameters of Laser Grating and Molecular Clusters a

Equations (7)

Equations on this page are rendered with MathJax. Learn more.

I ( x ) = ( 1 R ) I 0 + R [ 2 ( 1 C ) I 0 + 4 C I 0 cos 2 ( 2 π x λ ) ] .
t = 2 λ 0 λ / 2 exp [ n ( x ) ] d x = J 0 ( i n 0 2 ) exp [ n 0 ( 1 4 R C + 1 4 C ) ] J 0 ( i n 0 2 ) exp ( n 0 2 V λ ) .
Φ ( x ) = 4 π 2 α λ R C I 0 τ h c cos 2 ( 2 π x d ) Φ 0 ( 2 ) cos 2 ( 2 π x d ) ,
β = λ σ λ 8 π 2 α λ = n 0 ( i ) 2 Φ 0 ( i ) .
S ( Δ D ) = l = B l ( 1 ) ( 0 ) B 2 l ( 2 ) ( l T T T ) B l ( 3 ) ( 0 ) × exp [ 2 π i l d ( Δ D a T 2 ) ] ,
B n ( i ) ( ξ ) = exp ( n 0 ( i ) / 2 ) ( sin π ξ β cos π ξ sin π ξ + β cos π ξ ) n / 2 × J n [ sgn ( 1 β sin π ξ + cos π ξ ) × n 0 ( i ) 2 1 β 2 sin 2 π ξ cos 2 π ξ ] ,
V = max x [ S ( x ) ] max x [ S ( x ) ] max x [ S ( x ) ] + max x [ S ( x ) ] [ 0,1 ] .
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.