• No results found

Quantum-to-classical transition in many-body bosonic interference

N/A
N/A
Protected

Academic year: 2021

Share "Quantum-to-classical transition in many-body bosonic interference"

Copied!
6
0
0

Bezig met laden.... (Bekijk nu de volledige tekst)

Hele tekst

(1)

J.J. Renema1)*, V. Shchesnovich2), R. Garcia-Patron3)

1) Complex Photonic Systems (COPS), Mesa+ Institute for Nanotechnology, University of Twente, PO Box 217, 7500 AE Enschede, The Netherlands,∗

2) Centro de Ciências Naturais e Humanas, Universidade Federal do ABC, Santo André, SP, 09210-170 Brazil. and 3) Centre for Quantum Information and Communication, Ecole Polytechnique de Bruxelles,

CP 165, Université Libre de Bruxelles, 1050 Brussels, Belgium.

Bosonic many-body systems are prominent candidates for a quantum advantage demonstration, with the most popular approaches being either a quantum simulation beyond the reach of current classical computers, or a demonstration of boson sampling. It is a crucial open problem to under-stand how resilient such quantum advantage demonstrations are to imperfections such as boson loss and particle distinguishability. We partially solve this problem by showing that imperfect multi-boson interference can be efficiently approximated as ideal interference of groups of smaller number of bosons, where the other particles interfere classically. Crucially, the number of bosons undergoing interference in our approxmation only depends on the level of imperfections, but is independent of the actual number of particles. This allows us to construct a simple but stringent benchmark for comparing many-body bosonic technological platforms.

In recent years, interest in quantum information pro-cessing has focused onto small-scale quantum computing machines, which could perform single tasks of scientific or technological interest faster than classical computers, and which can be constructed with current or near-future technology [1, 2]. An important milestone on the way to such a device is a demonstrtion of a quantum advantage, i.e. a problem at which a quantum machine convinc-ingly outperforms a classical computer [3]. Such prob-lems could either be of practical interest (e.g. in quan-tum chemistry) or specifically set up for the purpose of demonstrating a quantum advantage. The most promi-nent bosonic platforms for such a demonstration are cold atoms on optical lattices [4], ion traps [5] and integrated photonic circuits [6].

One promising candidate for the demonstration of a quantum advantage is boson sampling, where the task is to provide samples of the output of a system of many bosons undergoing interference. Aaronson and Arkhipov [7] provided strong evidence that this task cannot be sim-ulated efficiently (i.e. in polynomial time) on a classical computer. Boson sampling was initially proposed in the framework of linear quantum optics, but alternative im-plementations for ion traps [8], one-dimensional optical lattices [9] and superconducting qubits [10] already ex-ist. The presence of computational hardness means that as the system size increases, the classical computer is in-creasingly at a disadvantage. The point where a boson sampler is expected to outperform a supercomputer is around 50 bosons [11], an observation that has spurred a range of experimental efforts [12–18].

However, a major obstacle to demonstrating a quan-tum advantage is our lack of understanding of the impact of experimental imperfections. While for small imperfec-tions, boson sampling was shown to retain its hardness [16, 19–23], experimentalists typically have to deal with large imperfections. In that case, there is no a priori

reason why claims of computational hardness could be maintained.

One strategy to understand the impact of imperfec-tions is to construct classical algorithms which make use of an imperfection in a boson sampler to conduct an effi-cient simulation [25–28]. These algorithms work by dis-proof by counterexample: if any efficient classical simu-lation algorithm can be constructed for a given level of imperfections, that rules out boson sampling with that level of imperfections as a demonstration of quantum ad-vantage. These algorithms therefore serve to demarcate areas of the parameter space where a demonstration of a quantum advantage is impossible. Similar efforts are un-derway in qubit-based quantum supremacy demonstra-tions, where the role of imperfections is also an active topic of study [29–34].

In this work, we simultaneously address the role of the two major imperfections in bosonic many-body interfer-ence, namely distinguishability and linear particle loss. We achieve this goal by providing a classical simulation algorithm for many-body bosonic interference where each boson has a fixed probability η to be transmitted through the experiment and may also have some degree of distin-guishability x. This work extends our previous algorithm [28], which considered only distinguishability, to a com-bination of losses and distinguishability.

We find that any interference of n bosons with loss η < 1, where m bosons survive is well approximated by polynomially many ideal coherent interference processes of size k ≤ m, supplemented with m − k classical bosons. Remarkably, k only depends on scale-invariant parame-ters such η and x. Therefore, boson sampling with losses η is only as hard as computing a permanent of size k and it therefore serves no purpose to construct lossy boson samplers of size larger than k, i.e., k serves as quantifier of the real coherence of the boson sampler.

We find that a transmission of η > 0.88 is necessary to

(2)

Figure 1. A pictorial representation of a boson sampler with losses. We consider a system with n input particles and m particles at the output, where the particles are lost before they enter the interferometer. The symbols illustrate the notation used in this article: τ are all ways of selecting m particles from n, and σ(τ ) are all the ways of permuting those particles. In the figure, we give an example of both.

simulate boson sampling with 50 bosons at an accuracy level of 10%, and that the current best boson sampling platforms technologies are restricted to interference of 21 bosons under the same criterion. This shows that achiev-ing a demonstration of a quantum advantage through boson sampling requires more than the construction of high-rate, large-scale photonic systems, as was demon-strated previously [35]: it also requires a qualitative im-provement in the equipment used.

Our work answers the question of whether boson sam-pling with linear losses is classically simulable, which was identified as a major open problem in the field of boson sampling [23, 24, 40]. This result fills a gap between two previous results. First, Aaronson and Brod showed the classical hardness of an artificial loss model, in which a constant number of bosons is lost, irrespective of the number of input bosons [23]. Second, Oszmaniec and Brod, and Patron et al. showed clasical simulability of a loss model where the per-boson loss increases exponen-tially with the number of bosons [25, 26].

We begin by setting up the problem, see Figure 1. The initial boson sampling proposal concerns the interference of n single-boson input state over an N -mode coupling interferometer modeled by a unitary transformation U , acting on the annihilation operators. At the output we measure the particle number on each mode. A condi-tion for its hardness proof to hold is that the number of modes N obeys N = O n2, which guarantees that the probability of two bosons emerging from the same mode can be neglected. In that case, and without losses, the probability of the photons exiting the system in a par-ticular set of modes is given by P = |Perm(M )|2, where

M is a submatrix of the overall unitary transformation U that is chosen by selecting the rows and columns cor-responding to the input and output modes respectively, and Perm(M ) =P

σ

Q

iMσi,i is the permanent function

[36], where σ is a permutation and the sum runs over all

permutations.

The hardness of boson sampling ultimately stems from the fact that for an arbitrary matrix, the permanent can-not be approximated efficiently by a classical computer [37]. The best known algorithms for computing an arbi-trary permanent are due to Ryser and Glynn [38, 39]. These algorithms scale as p2p, where p is the size of

the matrix. It was shown recently [35] that by using a Metropolis algorithm, one can generate samples from a probability distribution where each entry is given by a permanent, at a cost of evaluating a constant number of permanents. The hardness of boson sampling therefore rests on the hardness of computing the individual output probabilities of a boson sampler.

In this paper, we study how this hardness is compro-mised by loss and distinguishability. If the losses in each path of the boson sampler are equal, which is a reasonable approximation in experiments, the action of the interfer-ometer is equivalent to a circuit where all losses act at the front of the experiment followed by an ideal interferome-ter M [23, 26, 35]. The stochastic nature of the losses will make m fluctuate according to a binomial distribution, but for the purpose of keeping the presentation simple we first present and algorithm for fixed pair m and n and return later to the analysis of the most general case. A second consideration is that this scenario fits a recently suggested proposal to circumvent the losses problem in boson sampling by enforcing specific combinations of n and m by post-selecting, as was done recently for m = 5, n = 7 [40]. Our result allows us to strongly constrain the viability of this post-selection approach to boson sam-pling.

Without loss of generality, we will consider the proba-bility P of an arbitrary collisionless output configuration. As we are assuming that only m bosons are detected out of the initial n, the detection probability at the output results from the incoherent sum over the n

m 

different ideal boson sampling terms,

P =X

τ

|Perm(Mτ)|2=

X

Pτ (1)

where τ is an m-combination of n and the sum runs over all such combinations. Each Pτ is the probability

cor-responding to lossless boson sampling if m bosons were injected in the modes τ .

Our strategy is to break up equation (1) into terms which correspond to classical transmission, two-boson in-terference, three-boson inin-terference, and so on. We will then show that boson losses reduce the weight of the high-boson interference terms, to such an extent that beyond some number k, which is only a function of η, these terms can be neglected. We can then use our approximation as the imput for a Metropolis sampler, which can sample ef-ficiently from our approximate probability distribution.

(3)

detection probability in the lossless, fully indistinguish-able case [41]: Pτ= |Perm(Mτ)|2= X σ Perm(Mτ,1∗ Mσ(τ ),1∗ ), (2)

where σ is a permutation of the elements of τ , the sum runs over all permutations, ∗ denotes the elementwise product, ∗ denotes complex conjugation, and Mσ,1

de-notes permuting the rows of matrix M according to σ and the columns according to the identity. We will use this notation throughout. For the purpose of keeping the mathematics simple, we shall derive our results for perfect wave function overlap between the bosons (i.e. perfect indistinguishability), reintroducing distinguisha-bility at the very end.

Equation 2 can be rewritten by grouping terms accord-ing to the number of fixed points (unpermuted elements) in each permutation σ [28]. When this is done, perma-nents of positive matrices arise, which can be approxi-mated efficiently [42]. Grouping terms by the size of the derangements j (i.e. the number of elements not corre-sponding to fixed points) and substituting equation (2) into equation (1), we have:

P = m X j=0 X τ X σj(τ ) X ρ Perm(Mτ,ρ∗ Mσ∗jp(τ ),ρ) × Perm(|Mσu j(τ ), ¯ρ| 2) (3) ≡ m X j=0 cj,

where we have made use of their independence to ex-change the outer two sums, and the notation σj(τ )

de-notes a permutation with m−j fixed points which is con-structed from the elements of τ , σjpis the permuted part of such a permutation σj, σju is the unpermuted part, ρ

is a j-combination of m, and ¯ρ is the complement of that combination. It should be noted that since the size of σjp grows with j, the sum over j serves to group terms by computational cost, from easiest to hardest [43]. Simul-taneously, one can interpret the j-th term as containing all interference processes involving precisely j bosons.

Our goal is to show that in equation (3), terms with large j carry less weight and can therefore be neglected beyond some threshold value k, which depends on the losses and the desired accuracy of the approximation. Since the j-th term represents quantum interference of j bosons, this amounts to showing that boson sampling with losses can be understood as boson sampling of some k < m number of bosons. k therefore defines the maxi-mum size of a boson sampler which can be constructed at a given level of losses and distinguishability.

In order to compute the error of truncating the outer sum of equation (3) at some k, we must understand the variance of each term. For the lossless case, we showed

previously that for large m and j, the variance is given by Var(cj) = m!2/N2m, which notably does not depend

on j [28]. We find that (see Supplemental Material) for the case with loss, the variance is given by:

Var(cj) =  n m   n − j m − j  m!2/N2m, (4)

where the factor n m



arises from the sum over τ , and

the factor  n − j m − j



arises from constructive interfer-ence between terms which have identical σp. Crucially, this factor is a decreasing function of j. This demon-strates the central physical intuition of our result: fluctu-ations at high boson number are dampened out when av-eraging over many possible combinations of input bosons. We can formalize this idea by computing the expected value of the L1-distance between our approximation ˜P

and the exact distribution P , i.e., D(P, ˜P ) =P

s|P (s) −

˜

P (s)|, where the sum runs over all collision-free outcomes s. As n and m go to infinity, maintaining the ratio m/n = η, In the Supplemental Material, we provide a derivation of the expected value of D(P, ˜P ) over the ensemble of Haar random unitaries

EU

h

D(P, ˜P )i≤qηk+1/(1 − η). (5)

We note that since we can only compute the expec-tation value of the distance our algorithm will fail for some fraction of unitaries. However, using a standard Markov inequality (see Supplemental Material) one can bound the probability that D(P, ˜P ) does not satisfy a given bound. If EU

h

D(P, ˜P )i ≤  one can shown that PhD(P, ˜P ) > /δiis upper-bounded by δ. We note that numerical simulations suggest (see Supplemental Mate-rial) that the scaling in δ might actually be much better than indicated by this bound.

Solving equation (4) for the maximum boson sampler size k gives the number of bosons beyond which our algo-rithm becomes efficient, given a user-defined probability of failure δ, error tolerance  and the value of η corre-sponding to the experimental setup, which scales as

k ≈ 2log 1/ + log 1/δ

log 1/η , (6)

which we note depends only logarithmically on  and δ. To use our results for sampling, we use our approxima-tion to equaapproxima-tion 3 as the input for a Metropolis sampler, that samples efficiently from our approximate probabil-ity distribution. The algorithm looks as follows: first, given a value of probability of failure δ, error tolerance , n and m, we compute the maximum boson sampler size k using equation (6). Second, randomly choose a set of candidate input and output modes, i.e. sample over

(4)

Figure 2. Maximum number of bosons k which can be in-terfered before our classical algorithm becomes efficient, as a function of the overall boson transmission η, including losses in the sources, interferometer and detectors. The horizontal dashed line indicates the point n = 50, which is convention-ally taken as the limit of a quantum advantage. The ver-tical dashed line indicates the corresponding requirement of η > 0.88 for E(D) = 0.1. Left inset: dependence of E(D) on η for k = {10, 20, 30, 40, 50}. Right inset : dependence of k on E(D) for η = {0.75, 0.8, 0.85, 0.9, 0.95}.

all possible choices of τ and M . Third, compute the ap-proximate output probability by evaluating equation (3) up to the k-th term for a fixed τ . Fourth, use this prob-ability to compute the acceptance ratio of a Metropolis sampler [35, 44]. Repeat steps 2-4 to generate more sam-ples. In order to compute this approximation, we need to evaluate permanents of size up to k, and we need to evaluate order m2k of them. [28].

The algorithm for simulating boson sampling with losses presented above can be used to rule out a quantum advantage in certain areas of the parameter space. Fig-ure 2 shows the restrictions which our algorithm places on losses: we show parametric plots solving equation (5) for k versus η, for  = 0.1, 0.01 and 0.001. We find that in order to have fifty-boson interference while maintaining an L1-distance of 0.1 with high probability, a

transmis-sion of η > 0.88 is necessary. The higher the desired accuracy of the classical algorithm, the higher k can be. Losses are not the single imperfection that a boson sampling device may suffer from. Experimental partial distingushibability of boson, which are in principle com-pletely indistinguishable particles, can have an impor-tant impact in the quality of an experiment. Following the treatment of [28], our algorithm leads to a very sim-ple treatment of both. We re-introduce a finite level of boson distinguishability. as the wave function overlap x = hψi|ψji for i 6= j, where |ψii is the wave function

of the i-th boson. As derived in the Supplementary Ma-terial, equation (5) still holds where α = ηx2 takes the

place of η. Regardless of the specific combination of η

Figure 3. Interchange between distinguishability parameter x and the transmission probability η, for E(d) = 0.1. The lines in this plot are contours of constant α = ηx2, for α corresponding to the number of photons indicated in the leg-end. The red shaded area in the top right is the region of the parameter space where our algorithm cannot rule out a quan-tum advantage demonstration. The black points indicate the values of η and x of various photon sources reported in the literature.

and x used to achieve it, our algorithm can approximate experiments with equal α equally well. Its value may therefore be taken as a figure of merit of the ability of an experiment to interfere large numbers of bosons.

Figure 3 shows the tradeoff between boson distin-guishability and loss which is implied by equation (5). The curves are plots of η = x2/α, where the α correspond

to a different values of the maximum boson sampler size k, as indicated in the legend. The black dots incidate var-ious photon sources reported in experiments [18, 45–50]. The plot was generated for  = 0.1. This plot therefore demarcates areas of the parameter space where interfer-ence of a given number of photons cannot be simulated at that level of accuracy. We note that today not even the best possible experiments meet the requirements for a scalable demonstration of quantum advantage: consider-ing the best interferometers (99% transmission) [40], the best detectors (93% efficiency) [51] and the best SPDC sources, we arrive at α = 0.755, which implies that any scalable boson sampling experiments with more than 21 photons will be simulable with our algorithm at the 10% accuracy level.

Finally, we consider proposals involving postselection on those cases where (almost) all bosons make it through, as is usually done in experiments. Wang et al. re-cently proposed such an experiment, generating 50 + p photons and detecting 50 [40]. Our results show than in such a case, assuming the distinguishability the photons produced by the quantum dots of that work remains at roughly x2 ≈ 0.95, p ≤ 3 is required in order not to be

(5)

level.

In real experiments, m would fluctuate according to a binomial distribution with mean ηn and variance η(1 − η)n. One can always efficiently simulate these fluctua-tions if one has an algorithm for fixed pair m and n, since there are approximately√m possible outcomes which oc-cur with high probability, and these are clustered around m = ηn. As shown in the Supplementary Material, its effect is only a small correction to equation (5) prefac-tor and a small correction to our empirical bound, which vanishes in the limit of large n. Therefore, the classical algorithm start by estimating a ˜k, given a value of prob-ability of failure δ, error tolerance , n. Then, we simply need to supplement the selection of the set of candidate output modes of the previously presented algorithm by a preliminary step that generates a random m.

Recently, an estimate of the loss which compromises the demonstration a quantum advantage was given by Neville et al. [35]. Our result improves on this result in several ways. First, we demonstrate how losses induce a transition from exponential to polynomial scaling in the number of photons, while their result is essential a runtime estimate comparing an inefficient classical calcu-lation against an inefficient experiment. Second, because our algorithm is polynomial in the number of particles, the bounds that we find are also much more stringent in an absolute sense. Third, we show that the required transmission is a monotonically increasing function of the number of bosons which is coherently interfered; whereas [35] was only able to show that lossy boson sampling is at most as hard as regular boson sampling, we show that it is in fact much easier.

Future improvement to this work could be a generaliza-tion to non uniform losses and more general disitinguisha-bility models and replacing the Metropolis sampler by a direct sampling algorithm, such as the one proposed by Clifford and Clifford [52] for exact boson sampling. An adaptation of this result to Gaussian boson sampling, an alternative approach for quantum supremacy that has ap-plication as a subroutine in a classical-quantum hybrid algorithm for the calculation of the vibronic spectra of molecules and finding dense subgraphs, would be also an interesting future research direction.

In conclusion, we have described an algorithm for sim-ulating boson sampling with linear loss. We have shown that this algorithm imposes a minimum requirement on the losses of an experimental setup required in order to demonstrate genuine interference of a given number of bosons. Remarkably, the threshold only depends on a scale-invariant parameter combining losses and indistin-guishability, which allows us to demarcate new areas of the parameter space where a quantum advantage is im-possible. This result will therefore serves as a guide to experimentalists aiming for such a demonstration.

Acknowledgements

We thank Pepijn Pinkse for critical reading of the manuscript. J.J.R. and R.G.P. acknowledge support from the Fondation Werner Anspach, and NWO Vici through Pepijn Pinkse. V.S. acknowledges CNPq of Brazil grant 304129/2015-1 . R.G.P. Acknowledges the support of F.R.S.- FNRS.

j.j.renema@utwente.nl

[1] R. de Wolf, Ethics Inf. Technol. 19, 271 (2017). [2] J. Preskill, arXiv:1801.00862 (2018).

[3] A.W. Harrow and A. Montanaro, Nature 549 7671, (2017).

[4] C. Gross and I. Bloch, Science 357 6355, (2017).: [5] R. Blatt and C.F. Roos Nat. Phys. 8 4:277-284, (2012). [6] F. Flamini, N. Spagnolo, F. Sciarrino, arXiv:1803.02790

(2018).

[7] S. Aaronson and A. Arkhipov, Theory Comput. 9, 143 (2013).

[8] C. Shen, Z. Zhang, and L.-M. Duan, Phys. Rev. Lett. 112, 050504 (2014).

[9] G. Muraleedharan, A. Miyake, and I.H. Deutsch, arXiv:1805.01858v2 (2018).

[10] S. Goldstein, et al Phys. Rev. B 95, 020502(R) (2017). [11] J. Wu et al., Nat. Sci. Rev. (2018).

[12] J. B. Spring et al., Science 339, 798 (2012). [13] M. A. Broome et al., Science 339, 794 (2012). [14] M. Tillmann et al., Nat. Photon. 7, 540 (2013). [15] A. Crespi et al., Nat. Photon. 7, 545 (2013). [16] M. Bentivegna et al., Sci. Adv. 1, e1400255 (2015). [17] J. Carolan et al., Science 349, 711 (2015).

[18] H. Wang et al., Nat. Photon. 11, 361 (2017). [19] V. S. Shchesnovich, Phys. Rev. A 89 (2014). [20] A. Arkhipov, Phys. Rev. A 92, 062326 (2015).

[21] S. Aaronson, Scattershot bosonsampling: a new approach to scalable bosonsampling experiments https://www.scottaaronson.com/blog/?p=1579. [22] A. P. Lund et al., Phys. Rev. Lett. 113, 100502 (2014). [23] S. Aaronson and D. J. Brod, Phys. Rev. A 93, 012335

(2016).

[24] S. Aaronson, Introducing some british people to p vs. np (comment 84) https://www.scottaaronson.com/blog/?p=2408.

[25] R. Garcia-Patron, J. J. Renema, and V. Shchesnovich, arXiv:1712.10037 .

[26] M. Oszmaniec and D. Brod, arXiv:1801.06166 .

[27] S. Rahimi-Keshari, T. C. Ralph, and C. M. Caves, Phys. Rev. X 6, 021039 (2016).

[28] J. J. Renema et al., Phys. Rev. Lett. 120, 220502 (2018). [29] J. Chen, F. Zhang, C. Huang, M. Newman, and Y. Shi,

arXiv:1805.01450 (2018).

[30] Z.-Y. Chen et al., Sci. Bult. 63, 964 (2018).

[31] M. J. Bremner, A. Montanaro, and D. J. Shepherd, Quantum 1, 8 (2017).

[32] E. Pednault et al., Breaking the 49-qubit barrier in the simulation of quantum circuits, 2017, arXiv:1710.05867. [33] S. Boixo et al., Nat. Phys. 10, 218 (2014).

(6)

[34] A. Deshpande, B. Fefferman, M. C. Tran, M. Foss-Feig, and A. V. Gorshkov, Phys. Rev. Lett. 121, 030501 (2018).

[35] A. Neville et al., Nat. Phys. 13, 1153 (2017).

[36] H. Minc, Permanents, Encyclopedia of Mathematics and its Applications. 6. (Addison-Wesley, 1978).

[37] L. Valiant, Theor. Comput. Sci. 8, 189 (1979).

[38] H. J. Ryser, Combinatorial Mathematics (Mathematical Association of America, 1963).

[39] D. Glynn, Eur. J. Comb. 31, 1887 (2010).

[40] H. Wang et al., Phys. Rev. Lett. 120, 230502 (2018). [41] M. C. Tichy, Phys. Rev. A 91, 022316 (2015).

[42] M. Jerrum, A. Sinclair, and E. Vigoda, A polynomial-time approximation algorithm for the permanent of a

ma-trix with non-negative entries (ACM Press, 2001). [43] P. P. Rohde, Phys. Rev. A 91, 012307 (2015). [44] L. Martino and V. Elvira, (2017), arXiv:1704.04629. [45] Sparrow quantum application note,

http://sparrowquantum.com/wp-content/uploads/2017/12/multiphotonapplicationnote.pdf. [46] Quandela edelight-r specifications

http://quandela.com/edelight.

[47] O. Gazzano et al., Nat. Commun. 4, 1425 (2013). [48] L. K. Shalm et al., Phys. Rev. Lett. 115, 250402 (2015). [49] M. Giustina et al., Phys. Rev. Lett. 115, 250401 (2015). [50] S. Slussarenko et al., Nat. Photonics 11, 700 (2017). [51] F. Marsili et al., Nat. Photonics 7, 210 (2013). [52] P. Clifford and R. Clifford, (2017) arXiv:1706.01260.

Referenties

GERELATEERDE DOCUMENTEN

We have calculated the excitation spectrum of an Andreev billiard in a magnetic field, both using a quasiclassical and a fully quantum mechanical approach. The quasiclassical

In addition to greatly simplifying the computation of chaotic behavior in dilute weakly interacting systems and providing a physical picture for the meaning of many-body chaos,

Determinant quantum Monte Carlo study of the screening of the one-body potential near a metal-insulator transition..

As applied to the bilayer problem, the novelty is that in any finite di- mension the classical theory becomes highly pathological: the bare coupling constant of the field

We have presented compelling numerical evidence for the validity of the theory of the Ehrenfest-time dependent sup- pression of shot noise in a ballistic chaotic system. 2,5 The

We use ihe open kicked rotator to model Lhe chaotic scatteimg in a balhstic quantum dot coupled by Iwo poinl contacts to election leseivoiis By calculating

Aangenomen wordt dat hooimijten, net als spiekers, belangrijke opslagplaatsen waren voor de oogst of wintervoer die (meestal) gelokaliseerd waren op het erf. 51 Daarnaast

Samenvattend over de jaren 20032005 kan worden geconcludeerd dat met alle variaties die de proef heeft opgeleverd, gemiddeld de gehanteerde CHO gehalten volgens het AspireNZ