vibrations, quanta and biology - arxiv · vibrations, quanta and biology ... keywords: biology,...

26
Vibrations, Quanta and Biology S. F. Huelga and M. B. Plenio a,b * a Institut f¨ ur Theoretische Physik, Universit¨at Ulm, Albert-Einstein-Allee 11, 89073 Ulm, Germany b Center for Integrated Quantum Science and Technologies, Albert-Einstein-Allee 11, 89073 Ulm, Germany Quantum biology is an emerging field of research that concerns itself with the experimental and theoretical exploration of non-trivial quantum phenomena in biological systems. In this tutorial overview we aim to bring out fundamental assumptions and questions in the field, identify basic design principles and develop a key underlying theme – the dynamics of quantum dynamical networks in the presence of an environment and the fruitful interplay that the two may enter. At the hand of three biological phenomena whose understanding is held to require quantum mechanical processes, namely excitation and charge transfer in photosynthetic complexes, magneto-reception in birds and the olfactory sense, we demonstrate that this underlying theme encompasses them all, thus suggesting its wider relevance as an archetypical framework for quantum biology. Keywords: Biology, Quantum dynamics, Environments, Vibrations, Excitons, Electrons, Protons, Transport, Coherence INTRODUCTION Following early speculations concerning the potential role of quantum physics in biology [1], recent progress in science and technology has led to the rapid emergence of a new direction of research whose aim is the experi- mental and theoretical exploration of quantum effects in biology (see e.g. [2, 3]) which are taking place on length and timescales that allow quantum dynamics and envi- ronmental fluctuations to enter an intricate and fruitful interplay. Before we enter into more detailed discussions, let us first make some points as to why the existence of quan- tum effects in biology may be considered surprising and why there is rapidly growing excitement for developing what is being called quantum biology. To begin with, biological systems are, almost by def- inition, open systems, as they need to be continuously supplied with energy to maintain the out of equilibrium state that life represents. Open systems, however, espe- cially warm, wet and noisy biological systems, are sub- ject to environmental fluctuations that are usually ex- pected to result in fast decoherence and, as a result, the suppression of well controlled quantum dynamics. Thus quantum phenomena may at first sight seem unlikely to play a significant role in biology. There are arguments however to counter this pessimistic view. At the level of molecular complexes and proteins, processes that are of fundamental importance for biological function can be very fast (taking place within picoseconds) and well lo- calised (extending across a few nanometers, the size of proteins) and may therefore exhibit quantum phenom- ena before the environment has had an opportunity to destroy them. Furthermore, early work in quantum in- formation science, for example, has shown that thermal noise in stationary non-equilibrium systems may in fact support the existence of quantum coherence and entan- glement [4, 5]. Hence the possible existence of significant quantum dynamics is not only a question of sufficiently short length and time scales but may also depend on a constructive interplay between a quantum dynamical sys- tem and its environment such that quantum correlations are not simply washed out or suppressed but may in fact be enhanced or regenerated by the interaction with the environment. These arguments suggest that quantum effects in bi- ology are possible at the right length- and time scales. Indeed, quantum phenomena such as electron tunneling [6, 7] have been observed in biological systems and there is some evidence for proton tunneling in enzymes (see e.g. [8]). As such, tunneling phenomena are not intimately related with biology. Electron tunneling for example is a well-known and important phenomenon in solid state physics. The question thus remains whether, on the one hand, biological systems will exhibit more complex quan- tum dynamical phenomena that may either involve sev- eral interacting particles or multiple interacting compo- nents of a network or, on the other hand, whether the specifics of the biological systems and their environments will play a crucial role in allowing or supporting certain quantum dynamical phenomena in biology. Only then would we call these ”non-trivial” quantum effects in bio- logical systems. We do expect however that in the course of evolution Nature will have learnt to make use of quantum phe- nomena only if these enable or make more efficient a useful biological function that provides an evolutionary advantage. It is indeed well established from a quan- tum information perspective that pure quantum dynam- ics of multi-component systems can provide qualitative performance improvements over classical systems for ex- ample where transport is concerned [9]. This provides arXiv:1307.3530v1 [physics.bio-ph] 27 Jun 2013

Upload: vantu

Post on 04-Jun-2018

220 views

Category:

Documents


0 download

TRANSCRIPT

Vibrations, Quanta and Biology

S. F. Huelga and M. B. Plenioa,b ∗

aInstitut fur Theoretische Physik, Universitat Ulm, Albert-Einstein-Allee 11, 89073 Ulm, GermanybCenter for Integrated Quantum Science and Technologies, Albert-Einstein-Allee 11, 89073 Ulm, Germany

Quantum biology is an emerging field of research that concerns itself with the experimental andtheoretical exploration of non-trivial quantum phenomena in biological systems. In this tutorialoverview we aim to bring out fundamental assumptions and questions in the field, identify basicdesign principles and develop a key underlying theme – the dynamics of quantum dynamicalnetworks in the presence of an environment and the fruitful interplay that the two may enter. Atthe hand of three biological phenomena whose understanding is held to require quantum mechanicalprocesses, namely excitation and charge transfer in photosynthetic complexes, magneto-receptionin birds and the olfactory sense, we demonstrate that this underlying theme encompasses them all,thus suggesting its wider relevance as an archetypical framework for quantum biology.

Keywords: Biology, Quantum dynamics, Environments, Vibrations, Excitons, Electrons, Protons,Transport, Coherence

INTRODUCTION

Following early speculations concerning the potentialrole of quantum physics in biology [1], recent progress inscience and technology has led to the rapid emergenceof a new direction of research whose aim is the experi-mental and theoretical exploration of quantum effects inbiology (see e.g. [2, 3]) which are taking place on lengthand timescales that allow quantum dynamics and envi-ronmental fluctuations to enter an intricate and fruitfulinterplay.

Before we enter into more detailed discussions, let usfirst make some points as to why the existence of quan-tum effects in biology may be considered surprising andwhy there is rapidly growing excitement for developingwhat is being called quantum biology.

To begin with, biological systems are, almost by def-inition, open systems, as they need to be continuouslysupplied with energy to maintain the out of equilibriumstate that life represents. Open systems, however, espe-cially warm, wet and noisy biological systems, are sub-ject to environmental fluctuations that are usually ex-pected to result in fast decoherence and, as a result, thesuppression of well controlled quantum dynamics. Thusquantum phenomena may at first sight seem unlikely toplay a significant role in biology. There are argumentshowever to counter this pessimistic view. At the levelof molecular complexes and proteins, processes that areof fundamental importance for biological function can bevery fast (taking place within picoseconds) and well lo-calised (extending across a few nanometers, the size ofproteins) and may therefore exhibit quantum phenom-ena before the environment has had an opportunity todestroy them. Furthermore, early work in quantum in-formation science, for example, has shown that thermalnoise in stationary non-equilibrium systems may in fact

support the existence of quantum coherence and entan-glement [4, 5]. Hence the possible existence of significantquantum dynamics is not only a question of sufficientlyshort length and time scales but may also depend on aconstructive interplay between a quantum dynamical sys-tem and its environment such that quantum correlationsare not simply washed out or suppressed but may in factbe enhanced or regenerated by the interaction with theenvironment.

These arguments suggest that quantum effects in bi-ology are possible at the right length- and time scales.Indeed, quantum phenomena such as electron tunneling[6, 7] have been observed in biological systems and thereis some evidence for proton tunneling in enzymes (see e.g.[8]). As such, tunneling phenomena are not intimatelyrelated with biology. Electron tunneling for example isa well-known and important phenomenon in solid statephysics. The question thus remains whether, on the onehand, biological systems will exhibit more complex quan-tum dynamical phenomena that may either involve sev-eral interacting particles or multiple interacting compo-nents of a network or, on the other hand, whether thespecifics of the biological systems and their environmentswill play a crucial role in allowing or supporting certainquantum dynamical phenomena in biology. Only thenwould we call these ”non-trivial” quantum effects in bio-logical systems.

We do expect however that in the course of evolutionNature will have learnt to make use of quantum phe-nomena only if these enable or make more efficient auseful biological function that provides an evolutionaryadvantage. It is indeed well established from a quan-tum information perspective that pure quantum dynam-ics of multi-component systems can provide qualitativeperformance improvements over classical systems for ex-ample where transport is concerned [9]. This provides

arX

iv:1

307.

3530

v1 [

phys

ics.

bio-

ph]

27

Jun

2013

2

further support for the expectation that nature has de-veloped non-trivial quantum phenomena in the dynam-ics of biological systems, possibly supported by their en-vironment whose presence and influence is unavoidable.These quantum phenomena are not merely a by-productof the underlying quantum nature of chemical bonds butare actually exploited by biological systems to enhanceperformance and achieve novel functionalities. The cleardemonstration that Nature makes use of quantum effectswould bring about the necessity for a significant change ofthinking for biologists as they would be required to graspquantum concepts in order to understand some funda-mental biological processes. The very same fact wouldhowever also present the opportunity to learn from biol-ogy by unraveling the mechanisms by which quantumdynamics and its interplay with environments lead toenhanced performance. The resulting design principleshave the potential to lead to the development of new ap-plications at the bio-nano scale [10].

Typical spatial scale [m]

10 -11 10 -10 10 -9 10 -8 10 -6 10 -5 10 -4 10 -2

10 -2

10 -3

10 -4

10 -6

10 -8

10 -12

10 -14

10 -1

Typ

ica

l ti

me

sca

le [

s]

Typical spatial scale [m]

10 -11 10 -10 10 -9 10 -8 10 -6 10 -5 10 -4 10 -2

10 -2

10 -3

10 -4

10 -6

10 -8

10 -12

10 -14

10 -1

Typ

ica

l ti

me

sca

le [

s]

Function

?

Tools

Quantum Classical

FIG. 1. Biological systems are organized in hierarchical struc-tures. The continuous refinement of experimental tools per-mits the investigation of ever finer detail giving rise to thediscovery of novel phenomena. At a certain level we expectquantum physical properties to become relevant. Whether na-ture has evolved to enhance them to take benefit from them(quantum enhanced efficiency) or to suppress them to avoidtheir detrimental effects (quantum noise) represents one ofinteresting open question at the heart of quantum biology.(Figure courtesy of Alipasha Vaziri.)

Last but not least, the recent acceleration of the de-velopment of quantum biology also has a very practical,technological reason. Indeed, it is worthwhile noting thatquantum biology is not a new field but does go back along time, perhaps to Jordan’s book ”Die Physik unddas Geheimnis des Lebens” [1] in which he posed thequestion ”Sind die Gesetze der Atomphysik und Quan-tenphysik fur die Lebensvorgange von wesentlicher Be-deutung?” (Are the laws of atomic and quantum physics

of essential importance for life?) and coined the termQuanten-Biologie (quantum biology). So, why this re-newed and rapidly growing interest in the field? To un-derstand this, it is helpful to remember the developmentof quantum information science, whose theoretical foun-dations had been studied for some time by the 1990’s.This research had revealed that concepts of quantum in-formation have the potential to provide real performanceadvantages over classical systems. Crucially, however,it also became clear that quantum technologies had ad-vanced sufficiently to turn these theoretical ideas intoreality. This led to the emergence of a rapid develop-ment of both theory and experiment which is continuingto this day.

In recent years, quantum biology too has been bene-fitting considerably from the refinement of experimentaltools that are beginning to provide direct access to theobservation of quantum dynamics in biological systems[11–14] thanks to their increasing sensitivity to quantumphenomena at short length and time scales (see fig. 1for a suggestive illustration of this point). These newlyfound technological capabilities have helped to elevatethe study of quantum biology from a largely theoreticalendeavor to a field in which theoretical questions, con-cepts and hypotheses may be tested experimentally andthus be subjected to experimental verification or falsifi-cation. Indeed we should stress here that experimentsare essential (even more so than in the well-controlledpresent day systems of quantum information science) toverify theoretical models because biological systems un-der investigation have a complexity and structural varietythat prevents us from knowing and controlling all theiraspects in detail. Results obtained by these refined ex-perimental techniques lead to new theoretical challengesand thus stimulate the development of novel theoreti-cal approaches. It is this mutually beneficial interplaybetween experiment and theory that promises an accel-erated development of the field.

In this tutorial overview we will explore the type ofphenomena that currently define the field of quantum bi-ology and aim to bring out what are key questions atthe present stage of development. We will then focus ourdiscussions on a principle that is rapidly gaining recog-nition as being of central importance in this field – thecrucial role of the interplay between quantum dynamicsof multi-site systems, a.k.a. networks, on the one handand of complex, structured environments on the other.We will illustrate the generality of this principle by usingit to analyse and interpret three key examples of quantumeffects in biological systems, environment assisted exci-tation energy transport in photosynthesis [11, 15, 16],magneto-reception of birds [17–19] and the mechanismunderlying olfaction [20, 21], for which there is somecompelling theoretical and/or experimental evidence thatquantum phenomena are essential for their understand-ing. These three examples will hopefully stimulate the

3

discovery of many more biological phenomena for whichnon-trivial quantum effects are of fundamental impor-tance and thus come to be seen as the seeds from whicha rich phenomenology of quantum effects in biology maygrow.

Protein

Molecules

Control of electronic and vibrational structures to

create & optimize function

Refined experiments

Classical Biology Quantum Biology

New numerical methods for modelling

Photosynthesis

Olfaction

Avian Magnetoreception

FIG. 2. Cartoon illustrating the broad line of argument inthis article: Following the identification of general questions,we will argue that biological systems use protein structure toadjust the properties of transport or sensory networks and, atthe same time, those of the environment of these networks.Mutual tuning of these structures through evolutionary adap-tation may achieve optimal performance which in turn can beexplained from generalizable design principles. These designprinciples can lead to the formulation of novel structures andexperiments to amplify and verify quantum effects. The ac-curate description of the interplay of structured environmentsand quantum dynamics especially in the non-perturbativeregime requires the development of novel theoretical meth-ods. These concepts will be discussed and shown to applyto the current three examples of quantum effects in biology,photosynthesis, avian magneto-reception and olfaction. It isthe hope that these examples will be joined by many othersand lead to the emergence of a new research branch, quantumbiology.

GENERAL QUESTIONS AND PRINCIPLES

In the following we would like to draw attention to anumber of general points related to the study of quantumeffects in biology and identify broad questions that mightbe worth exploring.

Biology, dynamics, transport — Although perhaps ob-vious, we would like to stress that biology is not merelyabout static structures but that the dynamics that is en-abled by these highly organized structures plays a keyrole. Indeed, this dynamics is essential as biological sys-tems need to be supplied with energy and drained ofentropy to maintain the out-of-equilibrium state that liferepresents. This requires a wide variety of transport pro-cesses including excitation energy transfer but extend-

ing to charge transport including electrons, protons andions as well as the transport of larger molecules, peptidesand proteins. But the role of transport processes extendsmuch further as they are also essential ingredients in pro-cesses such as signal recognition and transduction whichrequires the transfer of excitations or real physical parti-cles. Hence transport and more general dynamical phe-nomena play a fundamental role in biology and thereforein the following discussion of quantum effects in biology.

Quantum traits & advantages — Quantum physics of-fers a wide variety of features that distinguish it from itsclassical counterpart and which may in some cases allowfor enhanced performance and functionalities. This in-cludes coherence and interference, that is, wave-like fea-tures of particles which may lead to faster propagationfor example in quantum random walks [9]. More sophis-ticated multi-particle coherence phenomena distributedacross different components of a system such as entan-glement hold the potential to achieve higher sensitivityto external signals [22, 23]. Another quantum trait, thequantization of energy, leads to well-defined energy levelsthat can be excited in discrete portions, quanta of energy,only. In the exchange of energy between two quantumsystems these features will then enforce highest efficiencyif the respective energy levels, such as vibrational eigen-frequencies, are well matched. This in turn permits theunambiguous identification of frequencies and thus canfacilitate the construction of sensing devices [20].

Therefore elucidating to what extent and under whatconditions quantum traits may be realised in biologicalsystems and how they are exploited for enhanced perfor-mance is of considerable interest.

Open quantum systems — Biological structures in gen-eral are not isolated as they are open systems that are inpermanent contact with their environments. While manycombinations of system and environment are conceivable,in this text the system will tend to be formed of electronicdegrees of freedom (excitons, electrons, ...) while the en-vironment will either be of vibrational nature or com-posed of electron and nuclear spin degrees of freedom.While thermal fluctuations imparted by the environmentmay be beneficial for example to overcome potential bar-riers between distinct classical configurations and there-fore facilitating processes such as protein folding or chem-ical reactions, the benefit of system-environment inter-actions is less evident in the quantum world. The un-controlled fluctuations of these environments will lead tochanges in the local structures in which the electronicdegrees of freedom are embedded and may therefore leadto decoherence. The advantages of quantum coherencein random walks [9] may for example be destroyed bydecoherence [24] and therefore the normal expectationin quantum technologies tended to be that almost com-plete isolation of the quantum system from its environ-ment is required to reap the benefits of quantum effects.However, it was realized in that field that the interac-

4

tion between system and environment may lead to thecreation of quantum properties such as coherence andentanglement [4] and investigations in quantum biologyfound that they may enhance transport for example inphotosynthetic quantum networks [15, 16].

The elucidation of mechanisms by which the quan-tum dynamics on networks may enter a fruitful inter-play with their environment to achieve enhanced perfor-mance and long-lived quantum coherence in transport,signalling and sensing is thus of fundamental importance.

Environments are structured — As we shall discover,the role of the environment in biological systems is nota passive one. Biological environments are not feature-less sources of white noise nor do they represent merelyweak perturbations. The environmental spectral densi-ties [25] describing their interaction with the system tendto display two principal structures, a broad smooth back-ground which has a short memory time and interacts withthe system mainly through its fluctuations, i.e. causingnoise, and well defined narrow features corresponding, forexample, to long-lived vibrational motion that can lead toquasi-coherent dynamical non-equilibrium exchanges be-tween system and environment. As we will see later thesmooth background is due mostly to the protein environ-ment as well as noise processes originating from solventswhile the sharp features tend to originate from long-livedvibrations that belong to molecules that are held withinthe protein scaffold [26, 27].

The rich structure of biological environments leads tonon-Markovian dynamics which is also non-perturbativebeing neither weak nor strong compared to the intra-system dynamics. These features deviate from environ-ments that are typically studied in quantum technologiesand uncovering non-trivial consequences of these struc-tures that may have been exploited by nature and aretherefore worthy of careful study.

Out of equilibrium & back-action — Many processes ofinterest to quantum biology are initiated by the suddengeneration of an initial excitation, such as a quantum ofenergy to be transported, which drives the system awayfrom equilibrium. As a consequence of this sudden ex-citation neither the system nor its environment will re-main stationary. The ensuing out-of-equilibrium dynam-ics then leads to perturbation of the environment by thesystem and a back-action of the very same environmenton the system initiating as a result a dynamical exchangebetween system and environment. If the perturbation ofthe environment concerns the broad smooth backgroundit will relax back very rapidly to its equilibrium value dueto its short memory (correlation) time. If, on the otherhand, a narrow spectral feature, relating to a long-livedvibrational mode, is excited then this will lead to longlasting quasi-coherent motion of the environment withthe possibility of triggering quasi-coherent exchange withthe system. This and other non-equilibrium exchangesbetween system and environment can have considerable

influence on the system dynamics [28–35], influence thedirection of energy transfer [28] and may even lead to thegeneration of long-lived quantum coherence in the system[28, 36–38].

The functional role of this non-equilibrium exchangeand in particular the interaction of electronic system de-grees of freedom with long-lived vibrational modes rep-resents a key question in quantum biology. Thus the in-vestigation of the non-equilibrium, non-perturbative andnon-Markovian character of the system-environment in-teraction is of considerable interest. Besides the concep-tual impact that it may have it also calls for the devel-opment of novel methods for the accurate numerical de-scription of such system-environment interaction whichmust be capable of going well beyond the usual pertur-bative treatments in the description of most quantumtechnologies [87, 154, 163, 168, 178].

Control of Structure — The discussions so far sug-gest that quantum dynamics, structured environmentsand their mutual interplay may provide increased effi-ciencies and potentially even novel functionalities. Inorder to optimize this interplay between system and en-vironment, biological systems cannot simply be randomconglomerates but must possess structure that they cancontrol. Examples are proteins which are essentially one-dimensional amino-acid sequences, representing the pri-mary structure, that have the ability to fold into spe-cific three-dimensional arrangements whose short range,secondary, structure defines the local arrangement viathe relative orientation of the amino acid residues, whilethe global three-dimensional arrangement, denoted as thetertiary structure, provides the long-range structure. Itwill become crucial that far from being passive scaffoldswhose mere purpose is to hold other molecules in theirplace, the primary, secondary and especially the tertiarystructure of proteins play a much broader and active rolefor the emergence of quantum effects in biology. In fact,their purpose is to control and tune the properties of theactual molecular network, the structure of the environ-ment in which this network is embedded and, crucially,the interaction and interplay between the two. The pro-tein can achieve this by tuning properties such as localenergies, by adjusting the local environment mainly viaits secondary structure, and the coupling rates within thetransport network, by adjusting the distance and relativeorientation of constituents as determined mainly throughits tertiary structure, and at the same time the propertiesof the environmental fluctuations that this transport net-works are subjected to are determined and tuned mainlyvia the secondary protein structure. The latter may notonly be achieved by structuring the protein itself but alsoby placing within it molecules that provide desirable vi-brational features. This optimization that is requiredhere will have taken place on evolutionary time scales.It is noteworthy though that, distinct from evolution-ary optimization, via the control of conformation within

5

an existing structure nature also possesses the means toswitch between different states of operation.

How this control is realised in detail and which config-urations exploit optimally quantum properties as well assystem-environment interaction is of considerable inter-est for the exploration of quantum effects in biology.

Technologies & Experiments: Hardware and Software– The questions and principles stated so far will guide ourthinking and allow theory to progress. It must be stressedhowever that the study of quantum effects in biology is anendeavour in which theory alone cannot succeed and hasto proceed hand-in-hand with experiment. So much isunknown about biological systems that assumptions thatare entering theoretical models as well as the predictionsof these models must be tested against experiment. In-deed, as we already stated in the introduction, the questfor quantum effects in biology has gained momentum re-cently in no small part due to the technological advancesthat are increasingly permitting the direct observationof quantum effects in biological systems and the prob-ing of these systems on previously unknown length- andtime-scales.

The future development of the field will depend criti-cally on the identification and development of technolo-gies, classical or quantum physical, that can lead todeeper insights into the workings of biological systemsat the length and time scales at which quantum effectscan act. New technologies and new experimental set-upsas well as protocols need to be developed (see e.g. [39–42]for a few examples). Furthermore, the complexity of thesystems under investigation suggests that methods fromfields such as signal processing [43–45] may prove fruitfulto optimize experiments and subsequent analysis.

It will be exciting to see how new experimental phe-nomena and theoretical principles will stimulate the de-velopment of novel experimental tools and clever proto-cols. This, we believe, will be essential for the devel-opment of the field as only irrefutable direct proof ofquantum phenomena and their functional role will forcebiologists to adopt quantum theoretical concepts to un-derstand foundational aspects of biology.

QUANTUM DYNAMICS IN BIOLOGICALENVIRONMENTS

Before we start to discuss a set of general and gener-alizable principles that govern the quantum dynamics ofbiological systems we would like to discuss here an exam-ple that presents us with a ”smoking gun” which suggeststhat the study of the interplay between quantum dynam-ics in the presence of structured environments may leadto interesting insights.

Reaction center

Antennae

FIG. 3. A schematic picture of the transport network suchas the one realized in the Fenna-Matthews-Olson complex.Molecules, such as Bacteriochlorophyll a (BChla), are ar-ranged in space giving rise to specific distances and relativeorientations between individual BChla thereby adjusting thestrength of the dipolar interaction (indicated by black arrows)between excitations on different BChla molecules. The sur-rounding protein (not shown) is also able to control the localenvironment of the BChla molecules and thus adjust their ex-citation energies. The specific arrangement and the nature ofthe environment determines the transport efficiency from theleft upper site (site 1 in the FMO) that accepts excitationsfrom the antenna complex and transfers them via the lowerright site (site 3 in the FMO complex) to the reaction centerwhere charge separation is initiated to begin the irreversiblyprocess of binding the exciton energy in chemical form.

Transport & Environment: An illustrative example

Let us consider a transport network that is composedof a set of highly absorptive entities such bacteriochloro-phyll molecules each of which may support an electronicexcitation (an Frenkel exciton). These molecules will bearranged in space by a protein scaffold and together theyform a pigment-protein complex. In our concrete exam-ple this will be the Fenna-Matthews-Olson (FMO) com-plex [46, 47] which forms an integral part of the photosyn-thetic light harvesting complexes of green sulphur bacte-ria [48]. Such complexes serve to transport electronicexcitations, excitons [49], in the presence of a vibrationalenvironment.

We will contend that nature can construct and opti-mize pigment-protein complexes, for example the FMOcomplex, to create transport networks that exploit boththe quantum dynamics of its electronic degrees of free-dom and their interaction with a structured vibrationalenvironment. The detailed mechanisms that nature hasused to achieve this and the design principles it is follow-ing will be discussed in the next section and then usedas the basis to understand other seemingly unrelated bi-ological quantum phenomena.

6

Let us first provide evidence, by means of a numeri-cal example, that such tuning and optimization may betaking place and that the vibrational environment mayindeed have a significant and indeed beneficial impact onthe transport dynamics of the FMO complex. Then wewill move on to derive generalizable conclusion from thesefindings. The full Hamiltonian describing the exciton-vibrational interaction as well as the exciton-exciton in-teraction is given by H = Hex +HI +HB where

Hex =

N∑n=1

En|n〉〈n|+1

2

∑m 6=n

[Jmn|m〉〈n|+ h.c.], (1)

HB =∑i,k

~ωka†ikaik, (2)

HI =1

2

∑n

[∑k

√Snkωk(ank + a†nk)|n〉〈n|+ h.c.].(3)

Here |n〉 describes an excitation on site n, the Jmn de-scribe the dipolar interaction between excitation on sitesm and n and the operators ank, a

†nk denote bosonic de-

struction and creation operators for the kth independentvibrational mode coupled to site n [46]. The exciton-mode interaction is determined by the strength of theirHuang-Rhys factors Snk [50]. Note, that HI describes apurely dephasing interaction because the vibrational de-grees of freedom have energies that are at least 10− 100times smaller than the excitation energies of the excitonsthus suppressing direct exciton-phonon interconversion.The dephasing can be understood to originate from thefact that vibrations will change the local environment ofeach site (e.g. moving charges) and thus affect the excita-tion energy of the relevant site [51]. We assume that thedynamics is dominated by contributions to the spectraldensity where each site interacts with its own indepen-dent environment an assumption that is corroborated byfirst-principles numerical studies of photosynthetic com-plexes [52, 54–57] and normal-mode analysis combinedwith quantum chemical methods [58, 59]. The interactionbetween site n and its environment is characterized bythe spectral density Jn(ω) =

∑k Snkω

2kδ(ω − ωk) which

is a joint property of the environment and the systemcombining the strength of interaction of modes with themode density.

The full dynamical equation also need to include twofurther contributions, one describing the spontaneous an-nihilation of an exciton and the concomitant loss of theenergy into the general environment at a rate γloss, aprocess that nature would like to avoid. Secondly, thetransfer of the excitation from a specific site, in the FMOcomplex, that is the site labeled 3, into the reaction cen-ter which is again described by an irreversible decay atrate γRC motivated by the fact that in the reaction cen-ter charge separation is achieved to irreversibly stabilizethe excitonic energy. Both contributions enter the equa-tions of motion via typical Lindblad terms [60] so that

we find the global time evolution of the density operatordescribing system and environment to be governed by

dρS,Edt

= −i[H, ρSE ]

−γlossN∑

n=1

(|n〉〈n|ρSE + ρSE |n〉〈n| − 2|0〉〈n|ρSE |n〉〈0|)

−γRC(|3〉〈3|ρSE + ρSE |3〉〈3| − 2|0〉〈3|ρSE |3〉〈0|) (4)

where |0〉 denotes the electronic ground state of thepigment-protein complex. Needless to say, these dynam-ical equations describing the full state of system and vi-brational environment are generally too complex to besolved exactly. Crucially however we will see shortly thatthe structure of the environment does play an importantrole and thus the development of methods that can cap-ture these features will be of considerable importance forquantitative studies of such systems. In order to bringout the main points clearly, we would like to delay the dis-cussion of these numerical challenges to final part of thisarticle. In order to obtain our first observation we followthe approximate treatments presented in [15, 16, 61, 62]and will improve on these later on.

To this end we will model the system-environment in-teraction perturbatively and derive a master equation forthe system evolution only to model the transport throughthe FMO complex under low light conditions, i.e. therate γin at which excitations enter the network via site1 is much smaller than the rate γRC at which excitationleave the network from site 3 into the reaction center.As a consequence, the mean population in the transportnetwork is much smaller than unity at any time. Thetypical but at first sight perhaps surprising result of suchan analysis is presented in Fig. 4. Contrary to what onemight have expected, the conductivity of the electronictransport network in the FMO complex (quantified as therate at which the reaction center is populated in steadystate divided by the rate at which excitations enter thetransport network) exhibits a maximum at a finite de-phasing rate, that is, dephasing noise can actually assistthe electronic transport [15, 16].

This immediately raises the question as to whether theregime that results in optimal transport performance isfound to be essentially classical in the sense that the dy-namics is well represented by a rate equation model orwhether, despite the dephasing noise, it remains firmlyin the quantum mechanical regime in which quantum co-herent dynamics is only weakly perturbed by dephasingnoise. Questions of this type can be answered both ina qualitative and a more quantitative manner. First, anexamination of the parameters that are typically enter-ing the dynamical equations when describing photosyn-thetic complexes reveals that the strength of the intra-system coupling, e.g. the dipolar interaction, is com-parable in strength to the system-environment coupling.Hence one already expects that the dynamics is taking

7

−2 −1 0 1 2 3 4 5 6

0.65

0.7

0.75

0.8

0.85

0.9

0.95

1

Log(Dephasing rate)

Co

nd

uct

ion

Eff

icie

ncy

FIG. 4. Plot of the conductivity of the FMO complex whereexcitations enter the FMO complex at site 1 and exit at site3 versus the overall strength of the dephasing noise due tothe interaction between electronic and vibrational degrees offreedom. The key observation is that optimal performance isfound for an intermediate level of dephasing noise. No noise orvery strong noise are counterproductive for the performanceof the transport network. See [62] for the Hamiltonain param-eters and the spontaneous emission rate, the noise is modeledas local dephasing Lindblad master equation in the site basis.

place in a regime in which neither dephasing noise norquantum coherent dynamics clearly dominate. This isfurther corroborated by the examination of the coherenceand entanglement [63, 64] properties of states [62, 65]and, more importantly, the dynamics of the system [66]which demonstrate quite clearly that on shorter length-and timescales quantum coherence is present in the sys-tems while for longer distances and times classical prop-erties dominate. This suggests that indeed, the optimaloperating regime in this setting is found to be ”halfway”between the classical and the quantum world.

In the light of recent discussions concerning the rel-evance of coherence in photosynthetic systems, we feelthat an important remark is in order here. The coherenceproperties of the states and dynamics of a system thatare observable in an experiment may depend very muchon the specific experimental set-up and the specific exci-tation regime that the system is subjected to. Excitationby incoherent sunlight as well as ensemble averages maysuppress the observed coherence and have tempted someresearchers to reach the conclusion that coherence maybe of no relevance in these systems under natural condi-tions. This however is not necessarily correct as the cru-cial point are the coherence properties of the underlyingdynamics and not of signatures of coherence in an exper-imental signal. It is the equations of motion that deter-mine the performance of each individual system and areunaffected by the nature of the initial preparation or en-semble averages that may obscure the observed coherence(consider a set of pendula each of which oscillates inde-

pendently from the others at a fixed frequency and phase.If the phase for each pendulum is chosen at random thenthe global signal appears incoherent while clearly eachpendulum is coherent). While natural conditions do notresemble laser light, laser spectroscopic experiments onindividual specimens provide the sharpest tools for theidentification of the dynamical equations that govern thesystem evolution and thus have a crucial role to play inthe determination of quantum effects in biological sys-tems.

These observations raise the questions as to why opti-mal performance is achieved in this intermediate regime.Answering this question will lead us to identify the dy-namical and structural principles that are underlying op-timal performance of quantum transport networks. Itwill drive us towards uncovering a rich interplay betweenelectronic degrees of freedom and their vibrational envi-ronment and point towards the possibility that naturehas optimized both electronic networks and vibrationalenvironment in an evolutionary process. This will be thesubject of the next section.

DESIGN PRINCIPLES

Here we will elucidate basic principles that have beenfound to underlie the fruitful interplay between vibra-tional environments and coherent quantum dynamics[15, 16, 28, 61, 62, 66, 67, 71–78]. Identifying and un-derstanding these principles at a deep, intuitive leveland seeing how nature may have used them to opti-mize performance does provide additional value even ifindividual processes in specific circumstances have beenknown in different physical situations. The principlesthat we present here will also allow us to bring under oneumbrella several seemingly unrelated biological phenom-ena, namely excitation, electron and proton transfer pro-cesses, the chemical compass model of magneto-receptionof birds and the mechanisms underlying olfaction, eachof which is suspected to be governed in an essential wayby quantum phenomena.

Controlling resonances – The phonon antenna

We will begin by elucidating a first principle which willprovide an understanding why optimal transport perfor-mance in the FMO complex can be achieved at inter-mediate noise levels. More importantly, this principleis sufficiently general to provide a mechanism that cansupport the surprisingly long-lasting oscillatory featuresobserved in recent ultra-fast laser spectroscopy experi-ments [11–14] and explain key aspects of the dynamicsthat may underlie the process olfaction.

We will approach this topic by means of a simple butinstructive question concerning the optimization of a sim-

8

ple transport network (see fig. 5 for a schematic repre-sentation of the following). Consider a network madeup of only three sites, namely site 1, which accepts exci-tations from the antenna and site 3 which is connectedto the reaction center. Both, site 1 and site 3 are fixedin their properties (position, orientation and excitationenergy). The system is completed by site 2 whose excita-tion energy, position and orientation, and hence dipolarinteraction strength with sites 1 and 3 we are free tochoose. We assume that site 3 provides the zero of ex-citation energy, site 1 has an excitation energy that is300 wavenumbers higher (for readers unaccustomed towavenumbers, note that 1 wavenumber, also denoted as1cm−1, corresponds to ω = 1.88 ·1011s−1). The questionthat we would like to answer concerns the optimal choiceof excitation energy, position and orientation of site 2 orin other words the optimal choice of the excitation energyof site 2 and its dipolar coupling strengths to sites 1 and3? As such, this question cannot be answered unambigu-

?

?

Antennae

Reaction center

FIG. 5. A network made up of only three sites, site 1 whichaccepts excitation from the antenna, site 3 which is connectedto the reaction center both of which are fixed in their prop-erties (position, orientation and excitation energy) as well asanother site 2 whose excitation energy, position and orienta-tion, and hence dipolar interaction strength with sites 1 and2 we are free to chose. What is the optimal choice of the ex-citation energy of site 2 and its dipolar coupling strengths tosites 1 and 3?

ously as we are missing a crucial piece of information,namely that of the structure of the environmental fluc-tuations. This structure is characterized by the spectraldensity of the environment which is a combination of thedensity of environmental modes and their individual cou-pling strength to the system. Typical spectral densitiesin pigment-protein complexes possess considerable struc-ture with sharp peaks originating from long-lived vibra-tional modes as well as a broad background whose maxi-mum tends to be in the range of around 200cm−1 whichwe will now assume for the subsequent optimization. Forthe sake of clarity, let us assume that the environmental

spectral density has a single maximum and thus takesroughly the shape depicted in figure 6. A numerical op-timization employing Redfield equations to take accountof the spectral structure of the environment (see [34, 61]for theoretical and numerical details) now finds that theoptimal position of site 2 is close to site 1 such that itexhibits a strong coherent dipolar interaction and closein excitation energy. Having found this numerical resultswe now would like to rationalize its origin and therebyarrive at a very useful design principle – the phonon an-tenna.

J2

)(SJ

0

J2

1

2

3

3

Dressed levels

1

2

FIG. 6. In the upper figure, two closely spaced energy levelsare separated from a third level to which excitations shouldbe delivered. They are subject to dephasing noise from an en-vironment with a finite bandwidth that exhibits a maximum.A coherent interaction between the upper two energy levelsleads to dressed states |±〉 with an energy splitting which,if matched to the maximum of the environment spectral den-sity, will optimize transport from the upper to the lower level.Hence the dressed states act as an antenna to harvest envi-ronmental fluctuations to enhance transport.

Indeed, the strong coherent dipolar interaction be-tween sites 1 and 2 suggests that we move to a newbasis made up of the eigenstates of the coherent partof the dynamics of these two sites, that is the exci-tonic states of that system or, for quantum opticians,the dressed state picture. This change of picture leadsus to rewrite the Hamiltonian eq. (3) that describes thesystem-environment interaction in the excitonic basis ofeigenstates {|en〉} of eq. (1), so that |i〉 =

∑n C

in|en〉,

and the coupling terms

HI =1

2

∑n,m

(Qnm|en〉〈em|+ h.c.), (5)

where

Qnm =∑ik

√SkωkC

inC

im(aik + a†ik). (6)

9

This leads us to two insights. Firstly, in the exciton(dressed state) basis the action of the dephasing noisenow leads to transitions between excitons, that is ampli-tude noise, which facilitates transport towards the lowerof the two exciton states. Secondly, the two excitons(dressed states) are separated by an energy differencethat is related to the coherent dipolar coupling strengthand the energy difference of sites 1 and 2. The dom-inant contribution to the transition between these ex-citons (dressed states) arises from those environmentalmodes whose frequency closely matches the energy dif-ference between dressed states. Indeed, the optimal so-lution is such that the energy separation of the dressedstates matches the maximum of the environmental spec-tral density, i.e. where the environmental fluctuationsare strongest, so that the environment may bring abouttransitions between the dressed states most effectively.In this sense, we can argue that the two eigenstates ofthe coupled Hamiltonian are tuned to harvest environ-mental fluctuations to achieve optimal excitation energytransport through the formation of a tunable ”phononantenna”.

Observing these two points alone, that is inducingstrong coherent coupling to move to a dressed basis andtuning the coupling such that it matches the maximumof the spectral density of the environmental fluctuations,one already comes close to the numerically obtained so-lution of the above optimization problem and can thusoptimize excitation energy transport. It should be notedthat the phonon antenna principle is also capable ofmaking predictions about more complex transport net-works such as that of the FMO complex. Indeed, it wasfound that the physically important relaxation pathwaybetween sites 1 and 3 is mediated by pigments whichare spectrally and spatially positioned by the protein toefficiently sample the spectral function of the proteinsfluctuations [34, 61]. Whether this optimality is a de-terminant in the emergence of this structure in nature isanother matter altogether, but it is striking how well thephonon antenna concept can be used to rationalise thesite energies and couplings of the pigments participatingin this pathway. The role of vibrational modes has alsobeen studied for other light harvesting complexes such asthe cryptophyte antenna protein phycoerythrin 545 [35].

It now becomes transparent why the optimal operat-ing regime for excitation energy transport may actuallybe found to be where coherent dipolar interactions andsystem-environment interactions, i.e. dephasing noise,are of broadly comparable strength. Indeed, if the en-vironmental noise is too weak, then the formation ofdressed states will present little benefit for transport. Onthe other hand if dephasing noise is very strong then itwill suppress the formation of the dressed states and thusof the phonon antenna effect in the first place. As a con-sequence, an intermediate regime where dephasing noiseof intermediate strength is present naturally appears as

the optimal operating regime according to the phononantenna. We will later see that there are a number ofadditional mechanisms in which noise and coherent dy-namics coexist to lead to similar conclusions [62, 79].

Before we move on to the discussion of these effects, itis instructive to highlight a connection that the phononantenna concepts shares with the concept of Hartmann-Hahn resonances [80] in nuclear magnetic resonance andspin sensing. Here one is faced with the challenge that thesensor, e.g., an electron spin, may not be resonant withan external system generating a signal (e.g., a differentelectron or nuclear spin). This problem may be overcomeby continuously driving the electron spin in the sensor ata strength that leads to dressed states whose splittingmatches the frequency of the external signal and thuspermits a strong coherent response of the sensor (transi-tions between upper and lower dressed states) [80–82].

With this in mind we now move on to bring out twoother conclusions and connections that we can draw fromthe phonon antenna concept in the case when the spec-tral density of the environment possesses very sharplypeaked features, that is well-defined long lived vibrationalmodes. Indeed, this will provide a both a mechanism ex-plaining the origin of long-lived coherences observed inrecent ultrafast spectroscopy experiments and a connec-tion to biological sensors.

Long-lived coherences as a non-equilibrium process

Experimental observations employing ultrafast 2-Dspectroscopy on various photosynthetic complexes exhib-ited long-lived oscillatory features which were interpretedas evidence for long-lived electronic coherence in the sys-tems under investigation [11, 13, 14]. Under this hy-pothesis electronic coherence appear to exhibit lifetimesthat can reach the picosecond range thus exceeding ex-pectations from condensed matter systems at least ten-fold. This interesting observation gave rise to a varietyof attempts for explanations of the long-lived coherencesincluding (i) overall reduction of dephasing [57, 83] whichare however not compatible with the observed very shortlifetimes of optical exciton coherences in the system, (ii)correlations in the noise sources between different sites[52, 53, 85] which are however not supported by firstprinciples calculations of spectral densities [54–57] andnormal-mode analysis combined with quantum chemi-cal methods [58, 59] and (iii) variations of the electronicstructure of the FMO complex [86] which are not suffi-cient however to explain the observed durations. In thefollowing we will show that the inclusion of significantcoupling of electronic motion to long-lived vibrationalmodes [28, 34] are capable of explaining the observations[36, 37] and even more so to give support to the ideathat vibrational motion plays and important role for elec-tronic transport, quantum or classical – a principle which

10

we will show here to be of broader importance in biology.

To gain an insight into possible mechanisms that sup-port long-lived electronic coherences in biological sys-tems, we will now take the phonon antenna principle toits extreme by applying it to a system-environment inter-action in which the broad features of the spectral densityare supplemented by sharp features due to the presence ofsome long-lived and well-defined vibrational mode as pro-posed in [28]. As we have learnt in the previous section,tuning the dipolar interaction in the electronic system tothe energy difference of the excitonic states that matchesthe maximum of the spectral density will maximize therate at which transitions between these states occur. Fora broad and smooth spectral density these transitions willbe dominantly incoherent (following essentially a Fermi’sgolden rule argument). In the presence of a long-lived vi-brational mode the phonon antenna principle remains thesame but, crucially, the nature of the interaction changesas the interaction between a single vibrational mode andthe electronic degrees of freedom is coherent, at least foras long as the mode itself remains coherent.

Let us examine this mechanism in more detail by con-sidering an exciton transport network in which excitonsenter in one site and exit in another. Initially the net-work is in its ground state and no excitons are present.This situation is depicted schematically in Fig. 7a wherea pendulum represents the long-lived vibrational modewhich we assume to be in a thermal state with small ex-citation number and where the black dot represents thepopulation of the ground state of the electronic system.The higher lying electronic levels, representing the vari-ous exciton eigenstates of the electronic system, are notexcited even at room temperature as the excitation en-ergy is in the range of eV. The initial (fast) injection ofan exciton, either coherently or incoherently, populatesone of the exciton states of the system (raised black dotin Fig. 7b) and creates a sudden force on the electronsand nuclei and thus change their equilibrium positions(compressed spring in Fig. 7b). Now the environmentwill start to react to these forces which initiates transientoscillations of the modes at approximately their naturalfrequency ωk. The continuous background of the spectraldensity will relax very rapidly into the new equilibriumstate as it contains a broad range of frequencies and thuspossesses a very short correlation time. The well-definedlong-lived vibrational mode will oscillate for a consider-able time (which can be up to several picoseconds) andwill interact with the electronic system (in Fig. 7c wesee that the spring connecting the vibrational mode tothe electronic motion is periodically expanded and com-pressed). This in turn leads to oscillations between dif-ferent exciton states. We make two observations: Firstly,these oscillations will have the largest amplitude betweenthose exciton states whose energy difference is nearly res-onant with the frequency of the vibrational mode (see

Fig. 7d). Secondly, we note that the sudden displace-

FIG. 7. Simplified mechanical illustration of a possible princi-ple behind long-lived coherence in biological systems. Detailsare presented in the main text.

ment of this mode implies that it is now found to be in adisplaced thermal state which will be close to a coherentstate if the frequency of the mode is such that its thermaloccupation number is low. As is well-known in quantumoptics a mode in a coherent state acts on a two-levelsystem essentially like a time-dependent classical drivingfield resulting in coherent Rabi-like oscillations followingapproximately the Hamiltonian

Hdriving ≈1

2

∑n 6=m

(〈Qnm〉(t)|en〉〈em|+ h.c.), (7)

where 〈Qnm〉(t) ∝∑

ik

√SkωkC

inC

im sin(ωkt) in the

above-mentioned approximation of the initial, transientand coherent response of the modes to exciton injection.As a result we will observe coherent transitions betweendissipative excitonic states and thus coherences that willlast for the coherence time of the vibrational mode –an expectation that can be corroborated by more so-phisticated numerical treatments employing methods de-scribed in [28, 87–89]. This electron-vibrational couplingregenerates electronic coherences in the system to replacethose that are continuously damped out by the fluctua-tions of the smooth background environment and mayeven lead to stationary entanglement [90]. We note thatin the language of chemical physics, this phenomenonarises through coherent non-adiabatic coupling (via dis-crete mode motion) which induces oscillatory crossing ofpotential surfaces. The physical picture presented hereillustrates a key point: electronic coherence may emergefrom transiently exciting robust, weakly dephasing vibra-tional coherences which then transfer back coherence tothose exciton transitions that are well matched to themode [28, 36, 37].

The importance of vibrational modes for interpret-ing experimental observations in multidimensional spec-

11

troscopy has only recently begun to be appreciated[28, 36–38, 87] and from the discussion above is seen tobe intimately related to the phonon antenna concept andin fact quantum sensing.

Phonon-assisted electron tunneling & olfaction

In the following we will explain how the coupling ofelectronic and vibrational motion may also underlie thefunction of biological sensors and exemplify these ideasat the hand of a mechanism suggested to be an importantcontribution to the function of olfactory sensors.

Despite considerable progress concerning the under-standing of the structure of olfactory receptors that in-volved in the early stages of the olfactory process, thedetailed mechanisms by which we are able to discrim-inate between the vast number of odorants are not yetfully understood [91]. This is emphasized by the fact thatfor nearly 100 years researchers have striven, with limitedsuccess, to identify principles that allow for the predic-tion of smell. The principal reason for this failure canbe traced back to the lack of a detailed understandingof what is actually happening during and shortly afterthe process of binding of odorants to the bindings site ofreceptors. There are currently two mechanisms proposedthat are based on quite different mechanisms and which,crucially, lead to different experimental predictions. It ispossible that both contribute in a critical way to olfac-tion.

On the one hand there is the idea of the lock-and-key principle. The olfactory stimulus is provided mymolecules, odorants, that arrive at the receptor via dif-fusion through the air. In the absence of the odorant thebinding pocket and the receptor exhibits small thermalfluctuations about some equilibrium conformation. Onlycertain types of odorants will be capable of attachingthemselves to the binding pocket, a choice determinedby chemical affinity, shape etc (this is the ”key in thelock” part of the principle). Once attached, the interac-tion between the odorant and the receptor results in achange of the average conformation of the receptor. Thisconformational change is then posited to induce furtherprocesses and initiates a signalling chain (this part con-stitutes the ”turning of the key” part of the principle).For many receptors, especially those binding only a veryspecific molecule, this appears to be a useful and validprinciple. Therefore it seems natural to adopt the verysame principle also for olfactory receptors. There are atleast 100000 odorants but far fewer olfactory receptors –several hundreds in humans, so that there is not a specificreceptor for each single odorant. The ability to differen-tiate such a large number of odorants would thus requirethat each odorants may bind to a variety of receptors.This would give rise to a vast number of distinct bindingpatterns and the subsequent sensation of smell.

Despite its attractiveness and applicability in certaincases the lock-and-key principle has to be subjected to ex-perimental test. This is where it has recently experiencedsome significant challenges. Firstly, it is not straightfor-ward to explain by means of the lock-and-key mechanismalone why outwardly very similar molecules may smellcompletely different while molecules with rather differentshape may smell similar. Even more remarkable is the ob-servation in recent experiments that Drosophila flies arecapable of discriminating between molecules in their hy-drogenated and their deuterated form and, importantly,are able to generalize from deuterated molecules to othermolecules that exhibit a vibrational modes similar in fre-quency to the Carbon-Deuterium stretch mode [21] (see[92] for recent experiments with similar outcomes on hu-mans).

Indeed, these observations provide some support for analternative theory that is based on physical properties ofmolecules rather than chemical or shape-based ideas thatare underlying the lock-and-key principle. Remarkably,it was well before the advent of the lock-and-key princi-ple that Dyson in 1938 proposed [93] that the smell ofa molecule may be determined by its vibrational spec-trum and that hence the olfactory system effectively op-erates as a vibrational spectrometer (and idea that hasalso been pursued later, starting in the 1950’s, by Wright[94]). This is an attractive idea, especially to a physicist,but at the time it suffered from a quite severe drawback –the lack of a plausible mechanism by which the olfactorysystem would in fact be able to identify specific vibra-tional modes. It was Luca Turin [20] who promoted theidea that inelastic electron tunneling (IET) may play arole in olfaction which, if true, would have the attrac-tive feature that it would help to predict the smell ofa molecule from the analysis of its vibrational spectrum[95]. IET is in fact a well established physical method fordetermining the vibrational spectrum of molecules [96].It was discovered in solid state physics by Jaklevik andLambe in 1966 when they identified anomalous behaviourin the conductivity through tunnel junctions in the pres-ence of organic adsorbates [97, 98]. They realised thatthese anomalies arose for certain voltages such that theenergy difference of the electrons on both sides of thetunneling barrier would match a vibrational quantum ofenergy of the adsorbed molecules and went on to demon-strate that this mechanism, inelastic electron tunnelingspectroscopy, allows them to identify vibrational spectraof molecules (apparently, they recognized its potential forthe theory of smell but were dissuaded by less far-sightedcolleagues [99]).

The proposed phonon assisted electron tunnelingmechanism for olfaction makes use of two phenomenawhich, together, make it a proper quantum biologicalphenomenon. There is on the one hand the tunnelingprocess of a massive particle, here the electron, and onthe other hand the fact that a vibrational mode, that is a

12

quantized harmonic oscillator, can only take up energy indiscrete quanta proportional to the relevant vibrationalfrequency ωodor. It is the second aspect that makes it pos-sible for this process to discriminate effectively betweendifferent vibrational modes and thus between the vibra-tional fingerprints of different molecules. A schematicpicture of the process is presented in Fig. 8. The bind-ing pocket of an olfactory receptor, which begins empty(upper part of Fig. 8), is assumed to be situated close toan electron donor and an electron acceptor (these mayeither be oriented alongside the binding site or separatedby the binding pocket itself, a difference that is impor-tant in practice but not relevant for the following basicargument). Crucially, it is assumed that electrons in thedonor have an energy that is ∆E higher than for electronsin the acceptor and that the energy distribution in donorand acceptor is much narrower than ∆E. In that casetunneling is suppressed in the absence of the odorant be-cause of the impossibility to satisfy energy conservation.If the energy difference ∆E is well matched to the ex-

D A

E

D A

E

FIG. 8. A schematic representation of the principle behindphonon assisted electron tunneling as a basis for olfaction. Abinding site is close to a electron donor and acceptor that areat different energies which suppresses tunneling. If a moleculewith a vibrational mode whose energy closely matches the en-ergy difference between donor and acceptor enters the bindingsite then electron tunneling can become energetically allowed.Then a charge redistribution becomes possible and a subse-quent signalling chain may be triggered.

citation energy of a specific vibrational mode ~ωodor ofthe odorant, then its presence may facilitate a detectableand very specific change in the dynamics of the receptorby the following process: The odorant enters the bindingpocket and structural changes may take place which givespace for the lock-and-key principle to play a certain role.Once the odorant has entered the binding pocket it will

reside there for a certain time long enough for electrontransfer to take place. The presence of the odorant witha vibrational mode that is well-matched to ∆E (see lowerpart of Fig. 8) then allows for energy conservation dur-ing a tunneling process – an electron can now tunnel fromdonor to acceptor while using its excess energy to excitea vibrational quantum of the odorant. This process willbe unidirectional for sufficiently low temperatures suchthat kT . ~ωodor. The concomitant charge redistribu-tion then triggers a further sequence of events that leadsto a signalling cascade.

Needless to say very detailed calculations are requiredto ascertain that the described mechanism is possible inprinciple for reasonable choices of parameters [100, 101].A wide variety of considerations need to be taken accountin these studies which include the determination of spec-tral densities describing the electron-vibration coupling,the effect that orientation of the odorant in the bind-ing pocket may have on it [102], temperature effects etc(see [103] for a more detailed analysis and [104] for anoverview).

What we would like to stress here though is the veryclose resemblance of the above mechanism to the phononantenna principle and thus the importance of electron-vibrational coupling. In both cases an electronic degreeof freedom senses the presence of a vibrational modedue to the tuning into resonance of the energy differ-ence of two states (dressed states in the phonon antennaand donor/acceptor in olfaction) to the energy of singlequanta of the vibrational motion. This resonance condi-tion ensures increased transport efficiency and thus leadsto a detectable effect at the physiological level (deliv-ery of an electronic excitation to the reaction center orelectron transfer which stimulates a subsequent responseof a receptor). Both examples demonstrate the potentialusefulness of controlled resonances for biological systems.

Noise assisted processes

Needless to say, not all biological systems will possesslong-lived vibrational modes that are strongly coupledto electronic degrees of freedom and can therefore play arole in the dynamics of the system. What is present inall biological systems however are thermal fluctuationsof the molecular and protein structures as well as thesurrounding solvents, water etc. These may lead to abroad noisy background mainly resulting in dephasingnoise on the electronic degrees of freedom.

Given the omnipresence of these noise sources one mayask as to whether nature has evolved to make use of inconjunction with quantum coherent dynamics to supportprocesses that are of relevance to life. That this mayindeed be the case we will explain in the following by firstpresenting several mechanisms by which dephasing noisemay in fact support transport phenomena. This will be

13

followed by a discussion of two phenomena in which noisecan be shown to assist fundamental biological processes.In the first example we will briefly revisit transport inphotosynthetic complexes while in the other example wewill show that the mechanism that is proposed to underliethe magneto-reception of birds depends crucially on thepresence of an environment.

Bridging energy gaps & blocking paths

Pigment-protein complexes consist of a number ofsites, schematically depicted in Fig. 3, whose excitationenergies generally exhibit a certain degree of static dis-order, that is, their on-site energies will differ from siteto site and also from one pigment-protein complexes toanother. If the energy difference between sites that ex-change excitation is larger than the intersite coupling ma-trix element in the relevant Hamiltonian, then transitionswill be severely reduced because of energy conservationunless of course there is are quasi-resonant vibrationalmodes present that, as was explained in the phonon-antenna mechanism, can take up the energy difference.The presence of a well-matched mode is not necessarythough. Broadband dephasing noise alone may alreadycome to the rescue in a manner that can be understoodfrom two different viewpoints. On the one hand one no-tices that dephasing noise will lead to a broadening of theexcitation energy of each site and thus to an increasedoverlap between the two energy levels while it does notcause the loss of excitations from the system (see Fig.9). Alternatively, one may take a dynamical viewpoint ofthe same phenomenon by realising that dephasing noisearises from the random fluctuations of the excitation en-ergies of each site. As a consequence, these fluctuatingenergy levels will occasionally come sufficiently close inenergy to allow for excitation energy transfer between thesites as the energy difference has been reduced to a valuesmaller than the direct coupling matrix element (see Fig.9). Again, we observe that this mechanism will lead to anoptimal operating regime at intermediate levels of envi-ronmental noise. Indeed, a low level of fluctuations willnot bring the site energies sufficiently close and trans-port remains suppressed, while excessive fluctuations ofthe site energies will reduce the time intervals in whichthe sites are energetically sufficiently close to allow forefficient energy transfer. This can be estimated easily bycomputing the overlap between Lorentzian lines of widthγ that are displaced by an amount ω0 in which case wefind

1

π2

∫ ∞−∞

γ

γ2 + ω2

γ

γ2 + (ω − ω0)2dω =

π(4γ2 + ω20). (8)

which takes on a maximum at γ = ω/2.An analogous behaviour arises already in the Marcus

theory of electron transport where the rate constant k

for electron transfer reaction is given by

k = Ae−−∆G

kBT , (9)

with ∆G given by

∆G =λ

4

(1 +

∆G0

λ

)2

(10)

which achieves a maximum for ∆G0 = λ. Here A is asystem dependent term, ∆G0 is the standard free energyof the electron transfer reaction and λ is the reorganiza-tion energy which quantifies the strength of interactionbetween the electron and its environment composed ofvibrational modes and the solvent [105].

deph

Site i

Site i Site j

Site j

deph

FIG. 9. Local dephasing, for example due to random fluctua-tions of the energy levels generated from random vibrationalmotion, leads to line-broadening and hence increased overlapbetween sites. Viewing these fluctuations dynamically, as il-lustrated by the double arrows, one finds that the energy gapbetween levels varies in time. The resulting nonlinear depen-dence of the transfer rate on the energy gap may thereforelead to an enhancement of the average transfer rate in thepresence of dephasing noise.

It is worth noting that while the application of an ex-cessive amount of dephasing noise suppresses transportthis may in itself serve a useful function if for example atransport network contains sites, for example for struc-tural reasons, that may lead to leakage of excitations intodomains from where further transport may be slow. Insuch a case dephasing may be useful to reduce the effec-tive transition rate to such sites and thus block unfavor-able transfer paths from being followed [74, 106].

Destructive interference, symmetry and noise

While linear networks can already exhibit interestingnoise assisted transport phenomena [16, 107–111], multi-site networks may exhibit more complex behaviour which

14

arises due to the interplay between a wealth of construc-tive and destructive interference effects in a quantum dy-namical system on the one hand and environmental noiseon the other hand [16, 62, 67–70, 77, 112].

A basic example that exhibits the essential nature ofthis type of effect consists of a simple three-site networkdepicted in Fig.10. Here two sites 1 and 2 are coupled to athird site which in turn leaks excitations irreversibly intoa reaction center. The coherent interaction is describedby a Hamiltonian

H =

3∑k=1

Ei|i〉〈i|+2∑

k=1

Jk3(|k〉〈3|+ h.c), (11)

where |i〉 corresponds to an excitation in site i and weassume J13 = J23. Let us begin by considering an exci-tation initially prepared in the antisymmetric state

|ψ〉 =1√2

(|1〉 − |2〉) (12)

which forms an eigenstate of this Hamiltonian whoseoverlap with the site 3, which we assume to be coupleddissipatively to a reaction center, vanishes. As a con-sequence this excitation will remain localized and doesnot propagate through the system. Eventually the finitelifetime of the excitation implies that it will be lost tothe general environment due to a spontaneous annihi-lation process, an event that is not in the interest of atransport network. Hence coherence effects may lead toa strongly reduced or even vanishing transport rate.

One may argue however, that under natural conditionsa pigment-protein complex is not excited in such an an-tisymmetric state, but will tend to receive a single exci-tation locally, for example on site 1 (this is for examplethe case of the FMO complex [46]). Nevertheless it iseasy to see that the subsequent dynamics has a propen-sity for leaving the system in an anti-symmetric state ormore generally a state that propagates slowly throughthe network due to quantum interference [62]. To thisend note that we can write the initial state localised onsite 1 as an equally weighted coherent superposition ofthe symmetric and the anti-symmetric states, i.e.

|1〉 =1√2

(|1〉 − |2〉√

2+|1〉+ |2〉)√

2

). (13)

Thanks to constructive interference a symmetric super-position experiences a coherently enhanced coupling tothe site 3 to which it will then propagate rapidly, andfrom there into the reaction center, while the antisym-metric part will not evolve at all. Hence, in 50% of thecases the system will remain in the anti-symmetric statewhile in the other 50% of the cases the excitation reachesthe reaction center. Therefore the transfer efficiency islimited to 50% in this setting.

Now it becomes evident that dephasing noise whosestrength is correctly tuned can have a beneficial effect

(a)

−12ψ

Site 3 +A+− −= AA

(b)

12ρ

Site 3

Site 2

Site 1

FIG. 10. A three-site network in which two sites 1 and 2are each coupled to a third site 3 via an exchange interactionof the same strength. Site three is irreversibly connected toa sink. In (a) the excitation is delocalized over two sites (redand green) with equal probability of being found at either sitebut with a wave function that is antisymmetric with respectto the interchange of red and green. This state will not evolvedue to destructive interference and hence no excitation willever reach the reaction center. In (b) pure dephasing causesthe loss of phase coherence and the two tunneling amplitudesno longer cancel, eventually leading to a complete excitationtransfer to the sink.

in such situations. Indeed, uncorrelated dephasing noiseacting locally on each site will randomly flip the rela-tive phase between |1〉 and |2〉 and thus leads to tran-sitions between the symmetric and the antisymmetricstate. Hence, the presence of dephasing noise inhibitsboth constructive and destructive interference and there-fore slows the propagation of an excitation in a systeminitiated in the symmetric state and accelerates the prop-agation of an excitation in a system initiated in the anti-symmetric state. As a consequence we expect again thatan intermediate noise level will be optimal, too low it willnot suppress destructive interference efficiently and toohigh it will suppress all transport. Noise will be benefi-cial if the overall propagation under the noisy environ-ment is still sufficiently rapid to be completed within thenatural lifetime of the excitation. Similar considerationshave also been conjectured independently to play a rolein biological electron transport in photosynthetic reac-tion centers [113] (note however that recent experimentaland theoretical work suggests that long-lived nuclear vi-brations are present in the reaction center dynamics andmay have a role to play here too [114–120]).

This is the simplest example for the more general phe-nomenon that complex quantum networks may possesssubspaces whose members do not propagate and do nothave overlap with the site connected to the reaction cen-ter. Systems in which such a subspace will be populatedduring the time evolution, e.g. because it includes thesite that receives energy, will benefit from environmentalnoise which can drive the systems out of the trappingsubspace [62]. It should also be noted that an energymismatch and the resulting time-evolution of the rela-

15

tive phase between the two sites 1 and 2 also leads totransitions between symmetric and anti-symmetric stateand can thus assists transport [16, 62, 121]. Which of thetwo processes if any plays a role in the actual dynamicsof the the FMO complex needs to be determined by care-fully designed experiments. In both cases the key pointis that some process breaks the symmetry of the systemand thus inhibits the system from getting trapped in anunfavorable state. It is this principle of breaking sym-metries that will also play a key role in another sensoryprocess in nature, namely the magnetic sense of birds aswe consider shortly.

Robustness of excitation energy transport in noisy quantumnetworks

As we have seen earlier in Fig. 4, noise does indeed sup-port transport through realistic transport networks suchas that of the FMO complex. We have identified a vari-ety of mechanisms for white as well as coloured noise tolead to these observations. Here we wish to highlight an-other aspect of noise assisted transport originating fromthe broad part of the spectrum, namely that it confers acertain stability on transport dynamics. In this case, nei-ther variations of the fine details of the spectral densitynor variations of the electronic structure of the transportnetwork have a significant effect. As shown in [62, 122]transport networks that exhibit efficient white noise as-sisted transport will suffer relatively small variations ofthe transport efficiency, in the sub-percent range, even ifthe electronic parameters of onsite and coupling energiesvary by up to 20%. Such a stability can be an attrac-tive feature to ensure robustness in an organism that issubject to changes in its environment.

Magneto-reception in birds

The effects of weak magnetic field on the growth ofplants as well as the remarkable orientation and navi-gation abilities of birds, mammals, reptiles, amphibians,fish, crustaceans and insects are well documented [123].The mechanism by which this magnetic field sense isachieved, however, is less well understood and at leasttwo alternative ideas are being considered, one that ex-ploits the forces exerted on ferrimagnetic iron oxide par-ticles embedded in the body, essentially a classical effect,and another that is based on magnetically sensitive freeradical reactions [17, 18, 124].

Behavioral experiments with birds such as the Euro-pean robin to study avian magneto-reception [125] haveled to the observation that the process of avian magneto-reception depends on the wavelength of the ambient light[126] and can be disrupted by very weak external oscil-lating magnetic fields [19, 127]. This together with the

experimental demonstration of magnetic field effects on aradical pair reaction at typical earth magnetic fields [128]provide pieces of evidence to support the idea that thechemical compass mechanism may be involved in avianmagneto-reception. The cryptochromes in the retina ofmigratory birds provide a potential physiological imple-mentation of such a mechanism [17] which has motivatedthe recently growth in attention from quantum physicists[124, 130–138].

Here we briefly outline the idea of the chemical com-pass based on the radical pair mechanism and stress thatit represents an example of the crucial role of the in-terplay between electronic spin quantum dynamics andthe nuclear spin environment of that electron spin (seeFig. 11 for a schematic representation of the following).A donor-acceptor pair is initially in its electronic groundstate characterized by a paired electron in a singlet state.Absorption of a photon induces an electron transfer ofa single electron from the donor to the acceptor thuscreating a radical pair, that is, two molecules with anunpaired electron each. For simplicity we assume thatthe electronic spin state remains unaffected in this stepso that the electrons remain in a singlet state. At thisstage no magnetic field sensitivity can be expected as thespin singlet state is rotationally symmetric and henceinsensitive to the orientation and magnitude of the ex-ternal magnetic field. Hence, a further ingredient is re-

Electron transfer

D + A

D

A

A

D

Singlet

Triplet

Incident Light

Magnetic Interaction

Singlet products

Triplet products

+

+

FIG. 11. A schematic description of the chemical compassof avian magneto-reception on the basis of the radical-pairmechanism. A photon excites two electrons from the groundstate to form a radical pair distributed across two molecules.The local hyperfine interaction leads to transitions betweensinglet and triplet space whose rate is affected by the orien-tation of the external magnetic field. Reaction products willdepend on the the electrons being in the singlet or tripletspace.

quired to break this symmetry. This ingredient is pro-vided by the nuclear spin environment of the donor and

16

acceptor molecules. Due to the distance dependence ofthe dipolar interaction between electron and nuclear spinthe unpaired electrons on donor and acceptor see domi-nantly uncorrelated interactions which induce symmetry-breaking transitions from the singlet to the triplet man-ifold. The Hamiltonian of the radical-pair plus nuclearenvironment is given by

H =∑

i=D,A

∑j

siTijIij − gµBB(sA + sD) (14)

where Iij and si are the nuclear and electron spin oper-ators respectively while Tij denotes the hyperfine cou-pling tensor. B denotes the external magnetic field, µ0

the Bohr magneton and g the gyromagnetic ratio. Thesetransitions and the resulting time varying population ofthe singlet and the triplet manifold will now be dependentof the relative orientation of the molecules with respectto the external magnetic field via the anisotropy of thehyperfine interaction as well as the strength of the mag-netic field as it dictates the energy splitting in the tripletmanifold.

Relative population differences between singlet andtriplet manifolds result in different rates of occurrence ofchemical products. A donor acceptor pair in the singletstate may either recombine to reach the groundstate orlead to a reaction product at rate kS . Spins in the tripletstate cannot recombine to reach the ground state but maylead to different reaction products at rate kT . Hence theamount and nature of reaction products will depend onthe relative populations of the singlet and triplet man-ifolds whose time evolution is usually described by asimple phenomenological quantum master equation (theHaberkorn approach [129])

d

dtρ = − i

~[H, ρ]−kS(QSρ+ρQS)−kT (QT ρ+ρQT ) (15)

where QS (QT ) are the projectors onto the singlet(triplet) subspace. The singlet yield is then determinedas kS

∫tr[ρ(t)QS ]dt.

The fact that local interactions between electron spinsand their nuclear environment lead to a breaking of thesymmetry of the electronic spin state gives a first indica-tion of the importance of the presence of local environ-ments. It does not however reveal the full depth of therole of the system-environment interaction. Indeed, con-trary to the assumption of early studies, it is not merelythe coherence properties of the electron spin state thatdetermine the sensitivity of the chemical compass. Togain a full understanding of the chemical compass andthe role of coherence one must study the entire systemcomposed of electron spins and nuclear spins [138]. Itturns out that it is the coherence of the initial state (elec-tronic singlet state and unpolarized nuclear spins) in thebasis made up of the eigenstates of the full system in theabsence of a magnetic field that has predictive power of

the sensitivity of the chemical compass. The principalreason is down to the fact that the coherences in sys-tem and environment that are present initially will accu-mulate phase factors due to the presence of an externalmagnetic field which in turn will be seen, when observingthe electron spin only, as oscillations between singlet andtriplet states. Indeed, a large amount of initial coherencein this picture correlates strongly with high sensitivity ofthe chemical compass and the impact of different typesof decoherence can be predicted from it. Hence it is onlywhen studying the properties of system and environmentthat we can gain and understanding of the chemical com-pass and an opportunity to predict optimal designs [138].

This example brings out quite clearly that it is the co-operation of electron spin quantum dynamics on the onehand and the interaction with the nuclear spin environ-ment of these electrons on the other hand that lead tothe magnetic field sensitivity of the radical pair mecha-nism. Without the environment no magnetic field sensi-tivity can be expected. Therefore a more refined pictureemerges when the structure of the nuclear spin environ-ment is taken account of by including it in the systemdynamics. From this viewpoint the chemical compasscan be seen as an environment assisted quantum inter-ferometer thus affirming avian magneto-reception as anexample of quantum biology that benefits from an inter-play of system-environment interaction [138].

Structure adaption

In the preceding section we have presented mechanismsby which the quantum dynamics of electronic systems canenter a beneficial interplay with the dynamics originatingfrom its interaction with an environment. The principlesexpounded here are more general though and apply toany quantum network whose task may include for exam-ple transport or sensing and that is in contact with anenvironment of charges, spins or vibrations. We have alsoexplained that it is likely that nature has optimized boththe properties of its quantum networks and the structureof the environment to achieve optimal performance.

Here we would like to summarize briefly some of themeans that nature has available for achieving this struc-tural adaptation and tuning for optimality. There aretwo principal aspects that can be controlled, the elec-tronic network and its environment. While in many re-spects it appears easier to achieve considerable changesin the electronic structure as compared to the vibrationalenvironment it is likely that both will have been subjectto evolutionary adaptation.

Position & Orientation: Interaction strength – Theprotein can arrange molecules, such as chromophores,controlling their respective distances as well as the rela-tive orientation of their dipole-moments. Thanks to the1/r3 distance dependence of the dipolar interaction and

17

its angular dependence 3(Di · rij)(Dj · rij) − (Di ·Dj),where Di is the optical dipole moment and rij the unitvector connecting the two molecules, significant changesof the interaction strength and even its sign can beachieved.

Static properties of Local Environments: Onsite exci-tation energies – In vacuum two realizations of the samemolecule, such as chlorophyll are identical in all its as-pects. In a protein scaffold however their properties canvary considerably. These variations are being controlledby the local environment such as the presence and dis-tance of partial charges or even free charges that willaffect the local excitation energy of a molecule throughtheir electrostatic interaction. These changes can be sig-nificant and can reach the range of 100′s of wavenumbersor more even for very small distance changes in the en-vironment. For larger structural changes of the proteinscaffold, e.g. due to variations in pH-value or other moredrastic changes the shift in wavelength can even reachdozens of nanometers. Thus the arrangement of the lo-cal environment is of crucial importance for the definitionof the electronic properties of a biological quantum net-work.

Dynamics of Local Environments: Dephasing noise –The structure of the local environment plays another cru-cial role as the motion of partial charges or even freecharges will lead to fluctuations in the excitation energyof the nearby sites. It is this process for example thatcouples the vibrational motion of proteins and moleculesto their electronic energies and is thus the cause of bothdephasing and the interaction of long-lived vibrationalmodes with electronic degrees of freedom.

Vibrational and spin environments – The spectral den-sity describing the interaction of electronic degrees offreedom and their environment is a combination of twoaspects, both of which may be tuned. Firstly, the cou-pling strength of individual environmental modes to thesystem which in turn depends on the amplitude of themotion of charges due to the vibration relative to thesite of interest and the magnitude and number of thesecharges. Secondly, the density of vibrational modeswhich is determined by the quasi-continuous mode den-sity of the protein as well as the discrete modes providedby the specific molecules that are realizing the quantumnetwork. For spin environments, e.g. the nuclear spinenvironment in magneto-reception, the structure of therelevant molecule and the form of the electronic wave-function determines the strength and anisotropy of thehyperfine interaction.

PC card principle – Besides the adjustment of thebroad vibrational spectrum of the protein a key principleis the ability to affect the discrete part of the vibrationalspectra by the choice of the molecule that realizes thequantum network. Each molecule will be characterizedby a set of vibrational modes of varying lifetime. De-pending on the fit of the molecule in the protein scaffold

the lifetime and coupling of these modes and thus theircontribution to the overall spectral density will be af-fected. In this way, it is conceivable that similar to aPC card the insertion of a specific molecule may allowfor certain functionality to be achieved. Likewise thephonon assisted tunneling process in olfaction suggeststhat this process may be used to recognise molecules andto initiate charge transfer processes.

Enhancing or decreasing noise – The presence of en-vironmental noise may or may not be of advantage for abiological system. Certainly, strong featureless noise canaffect the ability to act and react specifically to externalinput. Hence, in various circumstances it may be essen-tial to reduce the level of noise for example by protectinga molecule inside a protein structure or a hydrophobicbinding pocket. On the other hand it may on occasionbe beneficial to increase the level of noise by arrangingeasily movable charges close to a site which will respondstrongly to vibrations and thus create significant changesin the excitation energy of the site.

These are some of the principal means by which bio-logical systems may affect both their quantum dynam-ical networks as well as the environment around themto achieve optimal performance. This suggests that in-deed nature has at its hand a wide variety of possibilitiesfor tuning quantum networks and their environments toachieve robust and efficient devices.

NUMERICAL DESCRIPTION OF QUANTUMDYNAMICS IN STRUCTURED

ENVIRONMENTS

One of the key messages of the preceding sections is thespecial role of the interplay between the quantum dynam-ics of a system and its environment, in particular whenthis environment does not merely represent white noisebut possesses structure. Furthermore, it has becomeclear that the optimal operating regime for quantum biodynamics tends to favour parameter ranges in which theinteraction between system components is comparable tothe interaction of these system components and their en-vironment.

Both features are not well modeled by the traditionalperturbative treatments that lead to master equationsof Lindblad [139, 140], Redfield [141] or modified Red-field type [46, 142, 143] (see [60] for a review of masterequations). Indeed the essential importance of the non-Markovian nature of the system-environment interactioncalls for the development of non-perturbative methodsthat can accurately, certifiably and efficiently model theresulting dynamics. In recent years a wide variety ofmethods has been developed with the aim of address-ing some or all of these issues. In the following we willpresent the common theoretical setting in which thesemethods are formulated and then briefly outline key as-

18

pects of these methods. Details can be intricate and willbe left for further reading.

Standard model of open-quantum system

We begin by defining more formally a standard modelof an open quantum system as it is relevant to many ofthe biological settings that we have discussed so far. Inthis model the quantum system interacts with a macro-scopic number of environmental degrees of freedom andthe total system and environment state evolves under apurely unitary dynamics. This model is of considerableimportance for the description of a wide variety of biolog-ical environments. Dissipation and decoherence appearwhen the system is observed without any knowledge ofthe state of its environment, leading to a non-unitaryeffective dynamics for the sub-system’s reduced densitymatrix. The total Hamiltonian of system and environ-ment can be written as

H = HS +HE +HI , (16)

where HS is the Hamiltonian that refers to the degreesof freedom of the open quantum system of interest, HE

is the Hamiltonian describing the free evolution of theenvironment while HI describes the interaction betweenthe open quantum system and its environment.

The environment is assumed to be well described bya continuum of harmonic oscillator degrees of freedomlabeled by some real number. The internal dynamics ofthese harmonic oscillators is described by bosonic modesand is thus given by the Hamiltonian

HE =

∫ xmax

0

dx g(x)a†xax. (17)

In a physical context x would usually represent some con-tinuous real variable such as the frequency or momentumof each mode while xmax is the maximum value that thisvariable may assume. The creation and annihilation op-erators satisfy the continuum bosonic commutation rules[ax, a

†y] = δ(x−y) with the Dirac delta function δ(x−y).

We have adopted a continuum description of the envi-ronment but discrete environments can be treated in thesame way by replacing the integral by a summation.

The internal dynamics of the system is described by acompletely general Hamilton operator HS . For the in-teraction between the system and its environment weassume a linear coupling between arbitrary system op-erators and the position operator of the environment

HI =

∫ xmax

0

dxh(x)A(ax + a†x), (18)

where A is an arbitrary operator of the open quantumsystem and the (real) coupling function h(x) describes

the coupling strength with each mode. It is convenient,though not essential, to consider g(x) = x [144]. In thatcase the total Hamiltonian for system plus environmentis given by

H = HS +

∫ xmax

0

dx xa†xax +

∫ xmax

0

dxh(x)A(ax + a†x).

(19)For Gaussian initial states of the environment [145], thedynamics induced in the quantum system S by its inter-action with the environment is determined by the spec-tral density J(ω) [146, 147]. For the continuum modelof the reservoir that we are considering, i.e. choosingg(x) = x, this function is given by,

J(ω) = πh2(ω). (20)

The spectral function thus describes the overall strengthof the interaction of the system with the reservoir modesof frequency ω.

Despite the relative simplicity of this model its dy-namics is not exactly solvable and one has to resort tonumerical methods. This is particularly so, when theenvironmental spectral density does exhibit considerablestructure or when the coupling strength between systemand environment lies in the non-perturbative regime. Inthe following we discuss a variety of methods that havebeen developed in recent years to treat this situation.

Desiderata

There is a wide variety of simulations methods and inorder to evaluate them it will be helpful to first establishsome properties that one may wish a method to possess.Such desiderata for the method include (i) that it is ef-ficient in the system size, (ii) it is able to take accountof arbitrary spectral densities without losing efficiency orhaving to redesign the protocol, (iii) it has the ability todeal with both high, intermediate and low temperatures,and (iv) it should be certifiable, that is it possesses aknown and controllable error.

Time adaptive renormalization group methods

We begin by presenting an approach to the system-environment interaction that preserves the full informa-tion about the environment state, is able to treat ar-bitrary spectral densities and coupling strengths withina single unified framework with known and controllableerror and is computationally efficient in the size of theenvironment.

The key idea of this method which was developed in[87–89] is simple but effective. It begins by mapping thespin-boson model exactly onto a 1D system thus permit-ting the deployment of the time-adaptive density matrix

19

renormalization group (t-DMRG) technique to integratethe time evolution of the full system-environment dynam-ics efficiently.

We begin by demonstrating that a system linearly cou-pled to a reservoir characterized with a spectral densityJ(ω) is unitarily equivalent to a semi-infinite chain withonly nearest-neighbors interactions, where the system it-self only couples to the first site in the chain (see figure12) [148]. In other words, there exists a unitary oper-ator Un(x) such that the countably infinite set of newoperators

b†n =

∫ xmax

0

dxUn(x)a†x, a†x =∑n

Un(x)b†n (21)

satisfies the bosonic commutation relations [bn, b†m] =

δnm and leads to the transformed Hamiltonian

H ′ = HS+c0A(b0+b†0)+

∞∑n=0

ωnb†nbn+tn(b†n+1bn+b†nbn+1),

where c0, tn, ωn are real constants. The proof of this

)(xU n

bosonSpin

nnt

FIG. 12. Illustration of the effect of the transformation Un(x).On the left side a central system is interacting with a reservoirwith continuous bosonic or fermionic modes, parametrized byx, after Un(x) the system is the first place of a discrete semi-infinite chain parametrized by n.

statement becomes quite direct after one key observation.Since J(ω) is positive, we can always define the measuredµ(x) = h2(x)dx and write the unitary

Un(x) = h(x)pn(x) (22)

where pn(x) are some set of real orthonormal polynomi-als with respect to the measure dµ(x) = h2(x)dx withsupport on [0, xmax] [149]. Employing p0(x) = 1 we im-mediately find

HI = A(b0 + b†0).

HE in the new basis is obtained by employing the factthat orthogonal polynomials satisfy the recursion relation

pk+1(x) = (Ckx−Ak)pk(x)−Bkpk−1(x) for k = 0, 1, 2...and p−1(x) ≡ 0. We find

HE =∑n

[1

Cnb†nbn+1 +

An

Cnb†nbn +

Bn+1

Cn+1b†n+1bn]

with [88]

ωn = An/Cn, tn = Bn+1/Cn+1 = 1/Cn.

This approach provides us with a way to construct an ex-act mapping, one has just to look for a family of orthog-onal polynomials with respect to the measure dµ(x) =h2(x)dx. This can be done analytically in some impor-tant cases (see [88] for examples), however even if theweight h2(x) is a complicated function, families of or-thogonal polynomials can be found by using very sta-ble numerical algorithms such as the ORTHPOL package[150].

The reformulation of the system-environment interac-tion in the chain picture allows us to apply t-DMRGtechniques quite directly. As it can be proven for awide variety of spectral densities, in the large n limitthe chain parameters ωn to tn converge and their ra-tio approaches ωn/tn → 2. A translationally invariantharmonic chain with this ratio ωn/tn has a gapless dis-persion and excitations can escape down the chain with-out being scattered back towards the system (see Fig.13 for illustration). Thus the buildup of excitations inany region of the chain is limited over time, and corre-lations can be expected to be bounded. This in turnsuggests that t-DMRG will provide an accurate descrip-tion for long times [151]. The universal asymptoticsof environments revealed by this analysis are discussedin [88]. Practical experience, though not mathematicalproof, suggests that the method itself appears to satisfyall the desiderata listed above. Indeed it has been es-sential in the accurate numerical modeling dynamics ofcomplex environments which explain the emergence oflong-lived quantum coherence in photosynthetic complex[28] as explained earlier. It should be noted however that,in contrast to some other methods, it becomes computa-tionally less efficient with increasing temperature. On theother hand it easily incorporates time-dependent pertur-bations of the system, e.g. via laser pulses and it doespreserve all the information about the environment andis thus capable to following the dynamics after the exci-tation of complex wave-packets in the environment.

Hierarchy equations of motion

Recently, the numerical hierarchy technique [152–154],which has a longstanding history [155–157], has receivedrenewed attention in the context of excitation energytransfer across pigment-protein complexes. This ap-proach is non-perturbative and is capable of interpolat-ing, for example, between the Bloch-Redfield and the

20

(a)

(b)

FIG. 13. Illustrative sketch of open-system dynamics in thechain representation. (a) Subsystem initially injects excita-tions (shown as wave packets) into inhomogenous region ofthe chain. Scattering from inhomogeneity causes back ac-tion of excitations on the system at later times and leads tomemory effects and non-Markovian subsystem dynamics. (b)At long times, after multiple scattering, excitations penetrateinto the homogenous region and propagate away from thesystem without backscattering. This leads to irreversible andMarkovian excitation absorption by the environment.

Forster regimes [152, 154]. It derives a hierarchy of equa-tions in which the reduced density operator of the systemcouples to auxiliary operators which in principle allow forthe simulation of complex environments. The depth ofthis hierarchy and the structure of its coefficients dependon the correlation time of the bath and its spectral den-sity. This approach appears sufficiently flexible to takeaccount of spatial correlations in the noise as well. Forspecific choices of the bath spectral density, namely theBrownian harmonic oscillator and/or high temperaturesthe temporal bath correlations decay exponentially sothat the hierarchy can be terminated early with small butuncontrolled error and one obtains a manageable struc-ture of the hierarchy. An estimate suggests that in thiscase the set of operators in the hierarchy will scale atleast proportional τk where τ is the correlation time andk is the number of sites in the system to be studied (see[158] for a more detailed discussion). More complex spec-tral densities will lead to considerably more demandingevaluations of the coefficients of the hierarchy. A non-exponential decay of the temporal bath correlations alsoleads one to estimate an exponential growth of the num-ber of operators in the hierarchy. It is possible that spe-cific numerical simulations turn out to be more efficientthan those estimates suggest but there is no certificatethat provides error bounds. In fact, it is a challenge todetermine the errors introduced by the various approxi-mation steps that are involved in the numerical hierarchytechnique. Hence one has to test for convergence empir-ically by increasing the depth of the hierarchy until the

result does not change significantly any more. Neverthe-less, it should be noted, that the hierarchy method hasbeen applied successfully to a dimer [152, 154] as well asthe seven-site FMO complex [154, 159, 160] with a Brow-nian harmonic oscillator spectral density of the environ-ment. Note also recent numerically intensive numericalcalculations of 2-D spectra [161] using this method andrecent algorithmic improvements [162].

Path integral techniques

A variety of other methods to study transport pro-cesses and in consequence also transport in noisy environ-ments have been developed in condensed matter physics.These methods represent various approaches for findingnumerically the formal path integral solution of the timeevolution. These include the quasi-adiabatic path inte-gral approach [163–167] and the iterative summation ofreal-time path integrals [168] to name just two. Theseprocedures are expected to give good results in the hightemperature limit and hence short correlation time ofthe environments. With decreasing temperatures andthus increasing correlation time the computational ef-fort grows rapidly. Temperatures not too far below thetypical system frequencies appear to be accessible [169].For highly structured environments in which for exam-ple both narrow and broad features are combined thesemethods find challenges. In this case, sharply peakedmodes can be added to the system and their damp-ing is treated in the bath [170] but such an approachwill be challenging for the treatment of quantum net-works when addition of modes make the network itselftoo high-dimensional for numerical treatment. Path in-tegral methods have been applied mostly to dimers (seee.g.[166]) and their scaling to larger systems, just as forthe transformation approach to be discussed below, re-mains to be demonstrated. A comparison of the hierar-chy and the path integral methods has been carried outrecently for a Brownian harmonic oscillator environmentfor which both methods are expected to be applicable[171].

Other numerical methods

Many more methods exist which include methodsbased on a combination of the polaron transformationand a variational ansatz [172–174], schemes based on lin-earization of the environment dynamics [175–178], semi-classical methods in which the interaction between thesystem and its vibrational environment are replaced bymean-field type terms thus obviating the need for ten-sor product structures [28], schemes based on time-convolutionless master equations [179, 180], methodsbased on quantum state diffusion [181, 182] . For a recent

Contemporary Physics 21

review see of simulation methods see [183].

SUMMARY & CONCLUSIONS

The preceding sections have identified and discussedquestions of interest to quantum biology. In this con-text the unavoidable presence of partially uncontrolledenvironments in biology, often perceived as noise, has ledus to recognize a theme of fundamental importance thatis central to the study of quantum effects in biology –the interplay between quantum coherent dynamics of asystem on the one hand and the interaction of this verysystem with its environment on the other.

It is this unavoidable lack of isolation that provides theboundary conditions under which natural evolution hadto operate and therefore it may, in hindsight, not be sur-prising that nature has found solutions in which optimalbiological quantum dynamics tends to be achieved in aregime where the interaction within the quantum systemare of the order of its interaction with the environment sothat both contributions do not merely coexist but entera fruitful interplay.

That this is so, is not an accident. It can be understoodfrom simple and generalizable principles that we haveidentified in our discussions to explain that too muchor too little coherence can be detrimental and that it isin fact natural to expect that there is an intermediateregime that is optimal.

Moreover, we argued, biological systems have availablea variety of tools that enable them to achieve these op-timal regimes in a controlled fashion. Indeed, by the useof protein structure to arrange molecules and their localenvironment they are capable of tuning the propertiesof transport or sensory networks and, at the same time,adjust the environment of these networks by providingisolation or for example by inserting specific moleculeswith desirable vibrational or spin properties to fashionan environment. It is this toolbox that allows for mutualtuning through evolutionary adaptation which can thenbe used to achieve optimal performance whose origin weunderstand from generalizable design principles.

The importance of these design principles goes be-yond merely understanding what has been created al-ready but also paves the way by which these principles,when spelled out and made quantitative, can allow forthe rational design of optimal structures as well as theexecution of optimized experiments by which to amplifyand verify quantum effects in biology.

We have followed this path in the preceding sectionsand have determined and discussed such design princi-ples and have then applied them to bring under one um-brella the phenomena of photosynthesis, olfaction andavian magneto-reception. We hope that guided by theconsiderations and principles presented here we will beable to add to photosynthesis, avian magneto-reception

and olfaction, the three clouds at the otherwise blue skyof classical biology, and discover many more biologicalphenomena for which quantum effects are of fundamen-tal importance and thus come to be seen as the seedsfrom which a rich phenomenology of quantum effects inbiology may grow.

ACKNOWLEDGEMENTS

Over the course of the last few years we have benefittedfrom conversations and collaboration on topics of quan-tum biology with a wide variety of researchers, in partic-ular J. Almeida, A. Aspuru-Guzik, A. Bayat, J.M. Cai,T. Calarco, F. Caruso, F. Caycedo-Soler, A. W. Chin, A.Datta, M. del Rey, G. S. Engel, F. Jelezko, R. Ghosh, S.Montangero, A. Olaya-Castro, H. Plenio, J. Prior, Th.Renger, E. Romero, E. Solano, L. Turin, R. van Gron-delle and A. Vaziri. We gratefully acknowledge supportby the EU STREP project PAPETS, the ERC Synergygrant BioQ and an Alexander von Humboldt Professor-ship.

[1] P. Jordan, Die Physik und das Geheimnis des organis-chen Lebens, Friedrich Vieweg & Sohn, Braunschweig1943.

[2] M. Mohseni, Y. Omar, G.S. Engel, and M.B. Plenio(eds.), Quantum effects in biology, Cambridge Univer-sity Press, Cambridge 2013.

[3] Proceedings of the 22nd Solvay Conference in Chem-istry on ”Quantum Effects in Chemistry and Biology,Procedia Chemistry 3 (2011).

[4] M.B. Plenio and S.F. Huelga, Entangled light from whitenoise, Phys. Rev. Lett. 88 (2002), 197901.

[5] L. Hartmann, W. Dur and H.J. Briegel, Steady state en-tanglement in open and noisy quantum systems at hightemperature, Phys. Rev. A 74 (2006), 052304.

[6] D. DeVault and B. Chance, Studies of photosynthesisusing a pulsed laser I. Temperature dependence of cy-tochrome oxidation rate in Chromatium. Evidence fortunneling, Biophys J 6 (1966), pp. 825-847.

[7] D. DeVault, Quantum mechanical tunneling in biologicalsystems, Quart. Rev. Biophys. 13 (1980), pp. 387 - 564.

[8] Y. Cha, C.J. Murray and J.P. Klinman, Hydrogen tun-neling in enzyme-reactions, Science 243 (1989), pp. 1325- 1330.

[9] J. Kempe, Quantum random walks - an introductory re-view, Contemp. Phys. 44 (2003), pp. 307 - 327.

[10] G.D. Scholes, G.R. Fleming, A. Olaya-Castro, and R.van Grondelle, Lessons from nature about solar lightharvesting, Nature Chem. 3 (2011), pp. 763 - 774.

[11] G.S. Engel, T.R. Calhoun, E.L. Read, T.K. Ahn, T.Mancal, Y.C. Cheng, R.E. Blankenship, and G.R. Flem-ing, Evidence for wavelike energy transfer through quan-tum coherence in photosynthetic systems, Nature 446(2007), pp. 782 - 786.

Contemporary Physics 22

[12] I.P. Mercer, Y.C. El-Taha, N. Kajumba, J.P. Maran-gos, J.W.G. Tisch, M. Gabrielsen, R.J. Cogdell, E.Springate, and E. Turcu, Instantaneous mapping of co-herently coupled electronic transitions and energy trans-fers in a photosynthetic complex using angle-resolved co-herent optical wave-mixing, Phys. Rev. Lett. 102 (2009),057402.

[13] G. Panitchayangkoon, D. Hayes, K.A. Fransted, J.R.Caram, E. Harel, J.Z. Wen, R.E. Blankenship, andG.S. Engel, Long-lived quantum coherence in photosyn-thetic complexes at physiological temperature, Proc. Nat.Acad. Sci. Am. 107 (2010), pp. 12766 - 12770.

[14] E. Collini, C. Wong, K. Wilk, P. Curmi, P. Brumer, andG. Scholes, Coherently wired light-harvesting in photo-synthetic marine algae at ambient temperature, Nature463 (2010), pp. 644 - 649.

[15] M. Mohseni, P. Rebentrost, S. Lloyd, and A. Aspuru-Guzik, Environment-Assisted Quantum Walks in Pho-tosynthetic Energy Transfer, J. Chem. Phys. 129 (2008),174106.

[16] M.B. Plenio and S.F. Huelga, Dephasing assisted trans-port: Quantum networks and biomolecules, New J.Phys. 10, (2008), 113019.

[17] T. Ritz, S. Adem, and K. Schulten, A model forphotoreceptor-based magnetoreception in birds, Biophys.J. 78 (2000), pp. 707 - 718.

[18] K. Schulten, C. E. Swenberg, and A. Weller, A biomag-netic sensory mechanism based on magnetic field mod-ulated coherent electron spin motion, Z. Phys. Chem.NF111 (1978), pp. 1 - 5.

[19] T. Ritz, P. Thalau, J.B. Phillips, R. Wilschko, andW. Wilschko, Resonance effects indicate a radical-pairmechanism for avian magnetic compass, Nature 429(2004), pp. 177 - 180.

[20] L. Turin, A spectroscopic mechanism for primary olfac-tory reception, Chem. Sens. 21 (1996), pp. 773 - 791.

[21] M.I. Franco, L. Turin, A. Mershin, and E.M.C.Skoulakis, Molecular vibration-sensing component inDrosophila melanogaster olfaction, Proc. Natl. Acad.Sci. Am. 108 (2011), pp. 3797 - 3802.

[22] D.J. Wineland, J.J. Bollinger, W.M. Itano, F.L. Mooreand D.J. Heinzen, Spin squeezing and reduced quantumnoise in spectroscopy, Phys. Rev. A 46 (1992), pp. R6797- R6800.

[23] S.F. Huelga, C. Macchiavello, T. Pellizzari, A.K. Ekert,M.B. Plenio and J.I. Cirac, Improvement of frequencystandards with quantum entanglement, Phys. Rev. Lett.79 (1997), pp. 3865 - 3868.

[24] J.P. Keating, N. Linden, J.C.F. Matthews, and A. Win-ter, Localization and its consequences for quantum walkalgorithms and quantum communication, Phys. Rev. A76 (2007), 012315.

[25] The environmental spectral density is a combination ofmode density in the environment and the strength ofinteraction between the system and each of these modes[146].

[26] F. Caycedo-Soler, A.W. Chin, J. Almeida, S.F. Huelga,and M.B. Plenio, The nature of the low energy bandof the Fenna-Matthews-Olson complex: vibronic signa-tures, J. Chem. Phys. 136 (2012), 155102.

[27] M. Ratsep, Z.-L. Cai, J.R. Reimers and A. Freiberg,Demonstration and interpretation of significant asym-metry in the lowresolution and high-resolution Qy fluo-rescence and absorption spectra of bacteriochlorophyll a,

J. Chem. Phys. 134 (2011), 024506.[28] A. W. Chin, J. Prior, R. Rosenbach, F. Caycedo-Soler,

S. F. Huelga, and M. B. Plenio, Vibrational structuresand long-lasting electronic coherence, Nature Phys. 9(2013), pp. 113 - 118.

[29] J.D. Biggs and J.A. Cina, Studies of Impulsive Vi-brational Influence on Ultrafast Electronic ExcitationTransfer, J. Phys. Chem. A 116 (2012), pp. 1683 - 1693.

[30] G.H. Richards, K.E. Wilk, P.M.G. Curmi, H.M. Quiney,and J.A. Davis, Excited state coherent dynamics in light-harvesting complexes from photosynthetic marine algae,J. Phys. B 45 (2012), 154015.

[31] J.M. Womick and A.M. Moran, Exciton Coherence andEnergy Transport in the Light-Harvesting Dimers of Al-lophycocyanin, J. Phys. Chem. B 113 (2009), pp 15747- 15759.

[32] J.M. Womick and A.M. Moran, Vibronic Enhancementof Exciton Sizes and Energy Transport in Photosyn-thetic Complexes, J. Phys. Chem. B 115 (2011), pp. 1347- 1356.

[33] J.M. Womick, B.A. West, N.F. Scherer and A.M.Moran, Vibronic effects in the spectroscopy and dynam-ics of C-phycocyanin, J. Phys. B. 45 (2012), 154016.

[34] A.W. Chin, S.F. Huelga, and M.B. Plenio. Coherenceand Decoherence in Biological System: Principles ofNoise Assisted Transport and the origin of Long-livedCoherences. Phil. Trans. Roy. Soc. A 370 (2012), pp.3638 - 3657.

[35] A. Kolli, E.J. O’Reilly, G.D. Scholes, A. Olaya-Castro,The fundamental role of quantized vibrations in coherentlight harvesting by cryptophyte algae, J. Chem. Phys.137 (2012), 174109.

[36] V. Tiwari, W.K. Peters and D.M. Jonas, Electronicresonance with anticorrelated pigment vibrations drivesphotosynthetic energy transfer outside the adiabaticframework, PNAS 110 (2013), 1203 - 1208.

[37] J. Almeida, S.F. Huelga, and M.B. Plenio, Long-lived oscillatory signals in electronic 2-D spectra fromexciton-vibrational interaction, E-print arXiv:1307.xxxx

[38] N. Christensson, H.F. Kauffmann, T. Pullerits, and T.Mancal, Origin of long-lived coherences in light harvest-ing complexes, J. Phys. Chem B 116 (2012), pp. 7449 -7454.

[39] E. Harel, A.F. Fidler and G.S. Engel, Real-time mappingof electronic structure with single-shot two-dimensionalelectronic spectroscopy, Proc. Nat. Acad. Sci. 107(2010), pp. 16444 - 16447.

[40] F. Caruso, S.K. Saikin, E. Solano, S.F. Huelga, A.Aspuru-Guzik, and M.B. Plenio, Probing biologicallight-harvesting phenomena by optical cavities, Phys.Rev. B 85 (2012), 125424.

[41] F. Caruso, S. Montangero, T. Calarco, S.F. Huelga, andM.B. Plenio, Coherent open-loop optimal control of lightharvesting dynamics, Phys. Rev. A 85 (2012), 042331.

[42] S. Hoyer, F. Caruso, S. Montangero, Mohan Sarovar,T. Calarco, M.B. Plenio, and K.B. Whaley, Coher-ently controlled preparation and verification of excitonicstates in a light harvesting complex by ultrafast spec-troscopy with shaped pulses, arXiv:1306.xxxx

[43] E.J. Candes and M.B. Wakin, An Introduction To Com-pressive Sampling, IEEE Sig. Proc. Mag. 25 (2008), pp.21 - 30.

[44] J. Almeida, J. Prior and M.B. Plenio, Computation of2-D spectra assisted by compressed sampling, J. Phys.

Contemporary Physics 23

Chem. Lett. 3 (2012) 2692 - 2696.[45] J. N. Sanders, S. Mostame, S. K. Saikin, X. Andrade,

J. R. Widom, A. H. Marcus, A. Aspuru-Guzik, Com-pressed sensing for multidimensional electronic spec-troscopy experiments, J. Phys. Chem. Lett. 3 (2012)2697 - 2702.

[46] J. Adolphs and T. Renger, How proteins trigger excita-tion energy transfer in the FMO complex of green sulfurbacteria, Biophys. J. 91 (2006), pp. 2778 - 2797.

[47] M.S.A. Busch, F. Muh, M.E. Madjet, and T. Renger,The Eighth Bacteriochlorophyll Completes the Excita-tion Energy Funnel in the FMO Protein, J. Phys. Chem.Lett. 2 (2011), pp. 93 - 98.

[48] H. van Amerongen, L. Valkunas and R. van Grondelle,Photosynthetic Excitons, World Scientific 2000.

[49] A.S. Davydov, The Theory of Molecular Excitons, Sov.Phys. Usp. 7 (1964), pp. 145 - 178.

[50] M. Wendling, T. Pullerits, M.A. Przyjalgowski, S.I.E.Vulto, T.J. Aartsma, R. van Grondelle and H. vanAmerongen, Electron-vibrational coupling in the Fenna-Matthews-Olson complex of Prosthecochloris aestuariidetermined by temperature-dependent absorption andfluorescence line-narrowing measurements, J. Phys.Chem. B 104 (2000), pp. 5825 - 5831.

[51] V. May and O. Kuhn, Charge and Energy Transfer Dy-namics in Molecular Systems, Wiley-VCH Verlag 2004

[52] J. Strumpfer and K. Schulten, The effect of correlatedbath fluctuations on exciton transfer, J. Chem. Phys.134 (2011), pp. 095102.

[53] J. Lim, M. Tame, K.H. Yee, J.-S. Lee and J. Lee,Phonon-induced dynamic resonance energy transfer, E-print arXiv:1304.3967.

[54] C. Olbrich, J. Strumpfer, K. Schulten, and U.Kleinekathofer, Theory and Simulation of the Environ-mental Effects on FMO Electronic Transitions, J. Phys.Chem. Lett. 2 (2011), pp. 1771 - 1776.

[55] C. Olbrich, T.L.C. Jansen, J. Liebers, M. Aghtar,J. Strumpfer, K. Schulten, J. Knoester, and U.Kleinekathofer, From Atomistic Modeling to Excita-tion Transfer and Two-Dimensional Spectra of theFMO Light-Harvesting Complex, J. Phys. Chem. B 115(2011), pp. 8609 - 8621.

[56] C. Olbrich, J. Strumpfer, K. Schulten, and U.Kleinekathofer, Quest for Spatially Correlated Fluctu-ations in the FMO Light-Harvesting Complex, J. Phys.Chem. B 115 (2011), pp. 758 - 764.

[57] S. Shim, P. Rebentrost, S. Valleau, and A. Aspuru-Guzik, Atomistic study of the long-lived quantum coher-ences in the Fenna-Matthews-Olson complex, Biophys.J. 102 (2012), pp. 649-660.

[58] T. Renger, A. Klinger, F. Steinecker, M. Schmidt amBusch, J. Numata and F. Muh, Normal Mode Analy-sis of the Spectral Density of the Fenna-Matthews-OlsonLight-Harvesting Protein: How the Protein Dissipatesthe Excess Energy of Excitons, J. Phys. Chem. B 116(2012), pp. 14565 – 14580.

[59] T. Renger and F. Muh, Understanding photosyntheticlight-harvesting: a bottom up theoretical approach, Phys.Chem. Chem. Phys. 15 (2013), pp. 3348 - 3371.

[60] A. Rivas and S.F. Huelga, Open Quantum Systems – AnIntroduction, Springer Briefs in Physics 2012

[61] M. del Rey, A.W. Chin, S.F. Huelga, and M.B. Ple-nio, Exploiting structured environments for efficient en-ergy transfer: The phonon antenna mechanism, J. Phys.

Chem. Lett. 4 (2013), pp. 903 - 907.[62] F. Caruso, A.W. Chin, A. Datta, S.F. Huelga, and

M.B. Plenio, Highly efficient energy excitation transferin light-harvesting complexes: The fundamental role ofnoise-assisted transport , J. Chem. Phys. 131 (2009),105106.

[63] M.B. Plenio and V. Vedral, Teleportation, Entangle-ment and Thermodynamics in the Quantum World,Contemp. Phys. 39 (1998), pp. 431 - 446.

[64] M.B. Plenio and S. Virmani, An introduction to entan-glement measures, Quant. Inf. Comp. 7 (2007), pp. 1 -71.

[65] F. Fassioli and A. Olaya-Castro, Distribution of entan-glement in light-harvesting complexes and their quantumefficiency, New. J. Phys. 12 (2010), 085006.

[66] F. Caruso, A.W. Chin, A. Datta, S.F. Huelga, and M.B.Plenio, Entanglement and entangling power of the dy-namics in light-harvesting complexes, Phys. Rev. A 81(2010), 062346.

[67] J.S. Cao and R.J. Silbey, Optimization of Exciton Trap-ping in Energy Transfer Processes, J. Phys. Chem. A113 (2009), pp. 13825-13838.

[68] F. Caruso, S.F. Huelga and M.B. Plenio, Noise-Enhanced Classical and Quantum Capacities in Com-munication Networks, Phys. Rev. Lett. 105 (2010),190501.

[69] P. Giorda, S. Garnerone, P. Zanardi and S. Lloyd, Inter-play between coherence and decoherence in LHCII pho-tosynthetic complexes, E-print arXiv:1106.1986.

[70] J. Wu, R.J. Silbey and J. Cao, Generic Mechanism ofOptimal Energy Transfer Efficiency: A Scaling Theoryof the Mean First-Passage Time in Exciton Systems,Phys. Rev. Lett. 110 (2013), 200402.

[71] L. Campos Venuti and P. Zanardi, Excitation transferthrough open quantum networks: a few basic mecha-nisms, Phys. Rev. B 84 (2011), 134206.

[72] J.L Wu, F. Liu, Y. Shen, J.S. Cao and R.J. Silbey,Efficient energy transfer in light-harvesting systems, I:optimal temperature, reorganization energy and spatial-temporal correlations, New J. Phys. 12 (2010) 105012.

[73] S. Hoyer, M. Sarovar, and K.B. Whaley, Limits of quan-tum speedup in photosynthetic light harvesting, New J.Phys. 12 (2010), 065041.

[74] A.W. Chin, A. Datta, F. Caruso, S.F. Huelga, and M.B.Plenio, Noise-assisted energy transfer in quantum net-works and light-harvesting complexes, New J. Phys. 12(2010), 065002.

[75] P. Rebentrost, M. Mohseni, I. Kassal, S. Lloyd, and A.Aspuru-Guzik, Environment-Assisted Quantum Trans-port, New J. Phys. 11 (2009), 033003.

[76] J. Moix, J.L. Wu, P.F. Huo, D. Coker and J.S. Cao,Efficient Energy Transfer in Light-Harvesting Systems,III: The Influence of the Eighth Bacteriochlorophyll onthe Dynamics and Efficiency in FMO, J. Phys. Chem.Lett. 2 (2011), pp. 3045 - 3052.

[77] A. Olaya-Castro, C.F. Lee, F. Fassioli Olsen, andN.F. Johnson, Efficiency of energy transfer in a light-harvesting system under quantum coherence, Phys. Rev.B 78 (2008), 085115.

[78] M.B. Plenio and S.F. Huelga, Quantum dynamics of bio-molecular systems in noisy environments, Proceedingsof the 22nd Solvay Conference in Chemistry on ”Quan-tum Effects in Chemistry and Biology, Procedia Chem-istry 3 (2011), pp. 248 - 255.

Contemporary Physics 24

[79] S. Lloyd, M. Mohseni, A. Shabani and H. Rab-itz, The quantum Goldilocks effect: on the conver-gence of timescales in quantum transport, E-printarXiv:1111.4982

[80] S.R. Hartmann, and E.L. Hahn, Nuclear Double Res-onance in the Rotating Frame, Phys. Rev. 128 (1962),pp. 2042.

[81] J.-M. Cai, A. Retzker, F. Jelezko, and M.B. Plenio, ALarge-Scale Quantum Simulator on a Diamond Surfaceat Room Temperature, Nat. Phys. 9 (2013), pp. 168-173.

[82] J.-M. Cai, F. Jelezko, M.B. Plenio, and A. Retzker,Diamond-Based Single-Molecule Magnetic ResonanceSpectroscopy, New J. Phys. 15 (2013), pp. 013020.

[83] L.A. Pachon and P. Brumer, Physical Basis for Long-Lived Electronic Coherence in Photosynthetic Light-Harvesting Systems, J. Phys. Chem. Lett. 2 (2011), pp.2728 - 2732.

[84] Ch. Kreisbeck and T. Kramer, Long-Lived ElectronicCoherence in Dissipative Exciton-Dynamics of Light-Harvesting Complexes, J. Phys. Chem. Lett. 3 (2012),pp. 2828 - 2833.

[85] F. Fassioli, A. Nazir, and A. Olaya-Castro, QuantumState Tuning of Energy Transfer in a Correlated En-vironment, J. Phys. Chem. Lett. 1 (2010), pp. 2139 -2143.

[86] G. Ritschel, J. Roden, W. T. Strunz, A. Aspuru-Guzik,A. Eisfeld, Absence of quantum oscillations and depen-dence on site energies in electronic excitation transferin the Fenna-Matthews-Olson trimer, J. Phys. Chem.Lett. 2 (2011), pp. 2912 - 2917.

[87] J. Prior, A.W. Chin, S.F. Huelga, and M.B. Plenio, Ef-ficient simulation of strong system-environment inter-actions, Phys. Rev. Lett. 105 (2010), 050404.

[88] A.W. Chin, A. Rivas, S.F. Huelga, and M.B. Plenio, Ex-act mapping between system-reservoir quantum modelsand semi-infinite discrete chains using orthogonal poly-nomials, J. Math. Phys. 51 (2010), 092109.

[89] A.W. Chin, S.F. Huelga, and M.B. Plenio, Chain rep-resentations of open quantum systems and their nu-merical simulation with time-adapative density matrixrenormalisation group methods, in Semiconductors andSemimetals eds. Uli Wurfel, Michael Thorwart, EickeR. Weber and Chennupati Jagadish, vol 85, Burlington:Academic Press, 2011, pp. 115-144.

[90] S.F. Huelga, A. Rivas, and M.B. Plenio, Non-Markovianity assisted Steady State Entanglement, Phys.Rev. Lett. 108 (2012), 160402.

[91] M. Zarzo, The sense of smell: molecular basis of odorantrecognition, Biol. Rev. 82 (2007), pp. 455 - 479.

[92] S. Gane, D. Georganakis, K. Maniati, M. Vamvakias,N. Ragoussis, E.M.C. Skoulakis, and L. Turin, Molecu-lar Vibration-Sensing Component in Human Olfaction,PLOS ONE 8 (2013), e55780.

[93] G.M. Dyson, The scientific basis of odour, Chem. Ind.57 (1938), pp. 647 - 651.

[94] R. Wright, Odor and molecular vibration: neural codingof olfactory information, J. Theor. Biol. 64 (1977), pp.473 - 502.

[95] L. Turin, A method for the calculation of odor characterfrom molecular structure, J. Theor. Biol. 21 (2002), pp.367 - 385.

[96] J. Hihath and N. Tao, Electron-phonon interactions inatomic and molecular devices, Prog. Surf. Sci. 87 (2012),pp. 189 - 208.

[97] R.C. Jaklevic and J. Lambe, Molecular vibration spectraby electron tunneling, Phys. Rev. Lett. 17 (1966), pp.1139 - 1140.

[98] J. Lambe and R.C. Jaklevic, Molecular vibration spectraby inelastic electron tunneling, Phys. Rev. 165 (1968),pp. 821 - 832.

[99] L. Turin, The Secret of Scent. Adventures in Perfumeand the Science of Smell, Faber and Faber Ltd. London2006.

[100] J.C. Brookes, F. Hartoutsiou, A.P. Horsfield, and A.M.Stoneham, Could humans recognize odor by phonon as-sisted tunneling?, Phys. Rev. Lett. 98 (2007), 038101.

[101] I.A. Solovyov, P.-Y. Chang, and K. Schulten, Vibra-tionally assisted electron transfer mechanism of olfac-tion: myth or reality?, Phys. Chem. Chem. Phys. 14(2012), pp. 13861 - 13871.

[102] E.R. Bittner, A. Madalan, A. Czader, and G. Roman,Quantum origins of molecular recognition and olfactionin drosophila, J. Chem. Phys. 137 (2012), 22A551.

[103] J.C. Brookes, A.P. Horsfield, and A.M. Stoneham, TheSwipe Card Model of Odorant Recognition, Sensor 12(2012), pp. 15709 - 15749.

[104] J.C. Brookes, Olfaction: the physics of how smell works,Contemp. Phys. 52 (2011), pp. 385 - 402.

[105] R.A. Marcus, Electron transfer reactions in chemistry.Theory and experiment, Rev. Mod. Phys. 65 (1993), pp599 - 610.

[106] G.-Y. Chen, N. Lambert, C.-M. Li, Y.-N. Chen andF. Nori, Rerouting Excitation Transfer in the Fenna-Matthews-Olson Complex, E-print arXiv:1304.2613.

[107] K.M. Gaab and Ch. J. Bardeen, The effects of con-nectivity, coherence, and trapping on energy transfer insimple light-harvesting systems studied using the Haken-Strobl model with diagonal disorder, J. Chem. Phys. 121(2004), pp. 7813 - 7820.

[108] F.L. Semiao, K. Furuya and G.J. Milburn, Vibration-enhanced quantum transport, New J. Phys. 12 (2010),083033.

[109] A. Asadian, M. Tiersch, G.G. Guerreschi, J.M. Cai, S.Popescu, and H.J. Briegel, Motional effects on the ef-ficiency of excitation transfer, New J. Phys. 12 (2010),075019.

[110] A. Vaziri and M.B. Plenio, Quantum coherence in ionchannels: resonances, transport and verification, NewJ. Phys. 12 (2010), 085001.

[111] I. Kassal and A. Aspuru-Guzik, Environment-assistedquantum transport in ordered systems, New J. Phys. 14(2012), 053041.

[112] M. Mohseni, A. Shabani, S. Lloyd, Y. Omar and H. Ra-bitz, Geometrical effects on energy transfer in disorderedopen quantum systems, E-print arXiv:1212.6804.

[113] I.A. Balabin and J.N. Onuchic, Dynamically controlledprotein tunneling paths in photosynthetic reaction cen-ters, Science 290 (2000), pp. 114 - 117.

[114] M.H. Vos, F. Rappaport, J.C. Lambry, J. Breton, J.L.Martin, Visualization of coherent nuclear motion in amembrane-protein by femtosecond spectroscopy, Nature363 (1993), pp. 320 – 325.

[115] M.H. Vos, M.R. Jones, C.N. Hunter, J. Breton, J.L.Martin, Proc. Natl. Acad. Sci. 91 (1994), pp. 12701 –12705.

[116] A.M. Streltsov, A.G. Yakovlev, A.Y. Shkuropatov,V.A. Shuvalov, Femtosecond kinetics of electron trans-fer in the bacteriochlorophyll(M)-modified reaction cen-

Contemporary Physics 25

ters from Rhodobacter sphaeroides (R-26), Febs Letters,383 (1996), pp. 129 - 132.

[117] V.I. Novoderezhkin, A.G. Yakovlev, R. van Grondelleand V.A. Shuvalov, Coherent Nuclear and ElectronicDynamics in Primary Charge Separation in Photosyn-thetic Reaction Centers: A Redfield Theory Approach,J. Phys. Chem. B 108 (2004), pp. 7445 - 7457.

[118] S. Westenhoff, D. Palecek, P. Edlund, P. Smith and D.Zigmantas, Coherent picosecond exciton dynamics in aphotosynthetic reaction center, J. Am. Chem. Soc. 134(2012), pp. 16484-16487. (2012).

[119] E. Romero, R. Augulis, V.I. Novoderezhkin, M. Ferretti,J. Thieme, D. Zigmantas and R. van Grondelle, privatecommunication.

[120] C.A. Rozzi, S.M. Falke, N. Spallanzani, A. Rubio, E.Molinari, D. Brida, M. Maiuri, G. Cerullo, H. Schramm,J. Christoffers and Ch. Lienau, Quantum coherence con-trols the charge separation in a prototypical artificiallight-harvesting system, Nature. Comm. 4 (2013), 1602.

[121] J. Lim, J. Ryu, C. Lee, S. Yoo, H. Jeong, and J. Lee,Role of Energy-Level Mismatches in a Multi-PathwayComplex of Photosynthesis, New J. Phys. 13 (2011),103002.

[122] J.L. Wu, F. Liu, J. Ma, R.J. Silbey and J.S. Cao,Efficient energy transfer in light-harvesting systems:Quantum-classical comparison, flux network, and ro-bustness analysis, J. Chem. Phys. 137 (2012) 174111.

[123] S. Johnsen and K. J. Lohmann, The physics and neu-robiology of magnetoreception. Nature Rev. Neurosci. 6(2005), 703712.

[124] T. Ritz, Quantum effects in biology: Bird navigation,Procedia Chemistry 3 (2011), pp. 262 - 275.

[125] W. Wiltschko, R. Wiltschko, and T. Ritz, The mecha-nism of the avian magnetic compass, Procedia Chem-istry 3 (2011), pp. 276 - 284.

[126] J. B Phillips and S. C. Borland, Behavioural evidence foruse of a light-dependent magnetoreception mechanism bya vertebrate, Nature 359 (1992), pp. 142 - 144.

[127] P. Thalau, T. Ritz, K. Stapput, R. Wiltschko, and W.Wiltschko, Magnetic compass orientation of migratorybirds in the presence of a 1.315 MHz oscillating field,Naturwissenschaften 92 (2005), pp. 86 - 90.

[128] K. Maeda, K. B. Henbest, F. Cintolesi, I. Kuprov, C. T.Rodgers, P. A. Liddell, D. Gust, C. R. Timmel, and P.J. Hore, Chemical compass model of avian magnetore-ception, Nature 453 (2008), pp. 387 - 390.

[129] U.E. Steiner, T. Ulrich, Magnetic Field Effects in Chem-ical Kinetics and Related Phenomena, Chem. Rev. 89(1989), pp. 51

[130] I.K. Kominis, Quantum Zeno effect explains magnetic-sensitive radical-ion-pair reactions, Phys. Rev. E 80(2009), 056115.

[131] J.M. Cai, G.G. Guerreschi, and H.J. Briegel, Quantumcontrol and entanglement in a chemical compass, Phys.Rev. Lett. 104 (2010), 220502.

[132] E.M. Gauger, E. Rieper, J.J.L. Morton, S.C. Benjamin,and V. Vedral, Sustained Quantum Coherence and En-tanglement in the Avian Compass, Phys. Rev. Lett. 106(2011), 040503.

[133] J.M. Cai, Quantum Probe and Design for a Chemi-cal Compass with Magnetic Nanostructures, Phys. Rev.Lett. 106 (2011), 100501.

[134] J.M. Cai, F. Caruso, and M.B. Plenio, Quantum limitfor avian magnetoreception: How sensitive can a chem-

ical compass be?, Phys. Rev. A 85, (2012), 040304(R).[135] J. N. Bandyopadhyay, T. Paterek, and D. Kaszlikowski,

Quantum Coherence and Sensitivity of Avian Magne-toreception, Phys. Rev. Lett. 109 (2012), 110502.

[136] E. M. Gauger, S. C. Benjamin, Comment on QuantumCoherence and Sensitivity of Avian Magnetoreception,E-Print arXiv:1303.4539.

[137] H. J. Hogben, T. Biskup, and P. J. Hore, Entangle-ment and Sources of Magnetic Anisotropy in RadicalPair-Based Avian Magnetoreceptors, Phys. Rev. Lett.109 (2012), 220501.

[138] J.-M. Cai and M.B. Plenio, Chemical compass foravian magnetoreception as a quantum device, E-PrintarXiv:1304.4143.

[139] G. Lindblad, Generators of Quantum Dynamical Semi-groups, Comm. Math. Phys. 48 (1976), pp. 119 - 130.

[140] V. Gorini, A. Kossakowski, and E.C.G. Sudarshan,Completely Positive Dynamical Semigroups of N-LevelSystem, J. Math. Phys. 17 (1976), pp. 821 - 825.

[141] A. Redfield, On the theory of relaxation processes, IBMJ. Res. Dev. 1 (1957), pp. 19 - 31.

[142] W. M. Zhang, T. Meier, V. Chernyak and S. Mukamel,Exciton-migration and three-pulse femtosecond opticalspectroscopies of photosynthetic antenna complexes, J.Chem. Phys. 108 (1998), pp. 7763 7774.

[143] T. Renger and R. A. Marcus, Variable-range hoppingelectron transfer through disordered bridge states: Ap-plication to DNA, J. Chem. Phys. 107 (2003), pp. 84048419.

[144] In fact a redefinition of modes can always be made toachieve this choice with a suitable redefinition of h(x).

[145] These include thermal states and states displaced inphase space which covers a wide class of reasonable ini-tial states in biological systems. It will be highly chal-lenging to create states that differ significantly fromthese.

[146] A.J. Leggett, S. Chakravarty, A.T. Dorsey, M.P.A.Fisher, A. Garg, and W. Zwerger, Dynamics of the dis-sipative two-level system, Rev. Mod. Phys. 59 (1987),pp. 1 - 85.

[147] U. Weiss, Quantum Dissipative Systems, World Scien-tific, Singapore (2001).

[148] R.S. Burkey and C.D. Cantrell, Discretization in thequasi-continuum, J. Opt. Soc. Am. B 1 (1984), pp. 169- 175.

[149] P. Borwein and T. Erdelyi, Polynomials and PolynomialInequalities, Spinger 1995

[150] W. Gautschi, ORTHPOL-a package of routines for gen-erating orthogonal polynomials and Gauss-type quadra-ture rules, ACM Trans. Math. Soft. 20 (1994), pp. 21-62.

[151] J. Eisert, M. Cramer, and M.B. Plenio, Area laws forthe entanglement entropy - a review, Rev. Mod. Phys.82 (2010), pp. 277 - 306.

[152] A. Ishizaki and G.R. Fleming, Unified treatment ofquantum coherent and incoherent hopping dynamics inelectronic energy transfer: Reduced hierarchy equationapproach, J. Chem. Phys. 130 (2009), 234111.

[153] A. Ishizaki and Y. Tanimura, Quantum dynamics of sys-tem strongly coupled to low-temperature colored noisebath: Reduced hierarchy equations approach, J. Phys.Soc. Jpn. 74 (2005), pp. 3131 - 3134.

[154] A. Ishizaki and G.R. Fleming, Theoretical examina-tion of quantum coherence in a photosynthetic systemat physiological temperature, Proc. Nat. Acad. Sci. Am.

Contemporary Physics 26

106 (2009), pp. 17255 - 17260.[155] T. Takagahara, E. Hanamura, and R. Kubo, Stochastic

models of intermediate state interaction in 2nd orderoptical processes - stationary response .1, J. Phys. Soc.Jpn. 43 (1977), pp. 802 - 810.

[156] T. Takagahara, E. Hanamura, and R. Kubo, Stochasticmodels of intermediate state interaction in 2nd orderoptical processes - stationary response .2, J. Phys. Soc.Jpn. 43 (1977), pp. 811 - 816.

[157] Y. Tanimura and R. Kubo, Time evolution of a quantumsystem in contact with a nearly Gaussian-Markoffiannoise bath, J. Phys. Soc. Jap. 58 (1989), pp. 101 - 114.

[158] A. Ishizaki, T.R. Calhoun, G.S. Schlau-Cohen, and G.R.Fleming, Quantum coherence and its interplay with pro-tein environments in photosynthetic electronic energytransfer, Phys. Chem. Chem. Phys. 12 (2010), pp. 7319- 7337.

[159] A.G. Dykstra and Y. Tanimura, The role of the envi-ronment time scale in light-harvesting efficiency and co-herent oscillations, New J. Phys. 14 (2012), 073027.

[160] Ch. Kreisbeck, T. Kramer, M. Rodriguez, and B. Hein,High-performance solution of hierarchical equations ofmotions for studying energy-transfer in light-harvestingcomplexes, J. Chem. Theo. Comp. 7 (2011), pp. 2166 -2174.

[161] B. Hein, C. Kreisbeck, T. Kramer, and M. Rodrıguez,Modelling of Oscillations in Two-Dimensional Echo-Spectra of the Fenna-Matthews-Olson Complex, New J.Phys. 14 (2012), 023018.

[162] J. Zhu, S. Kais, P. Rebentrost, and A. Aspuru-Guzik,Modified Scaled Hierarchical Equation of Motion Ap-proach for the Study of Quantum Coherence in Photo-synthetic Complexes, J. Phys. Chem. B 115 (2011), pp.1531 - 1537.

[163] D. E. Makarov and N. Makri, Path -integrals for dissipa-tive systems by tensor multiplication - Condensed-phasequantum dynamics for arbitrarily long times, Chem.Phys. Lett. 221 (1994), pp. 482 - 491.

[164] N. Makri and D. E. Makarov, Tensor propagator foriterative quantum time evolution of reduced density-matrices .1. Theory, J. Chem. Phys. 102 (1995), pp.4600 - 4610.

[165] N. Makri and D. E. Makarov, Tensor propagator foriterative quantum time evolution of reduced density-matrices .2. Numerical methodology, J. Chem. Phys. 102(1995), pp. 4611 - 4618.

[166] M. Thorwart, J. Eckel, J.H. Reina, P. Nalbach, andS. Weiss, Enhanced quantum entanglement in the non-Markovian dynamics of biomolecular excitons, Chem.Phys. Lett. 478 (2009), pp. 234 - 237.

[167] P. Nalbach and M. Thorwart, Quantum Coherence andEntanglement in Photosynthetic Light-Harvesting Com-plexes, in Quantum efficiency in complex systems, PtI: Biomolecular systems, Book Series: Semiconductorsand Semimetals, Vol 83 (2010), pp. 39 - 75.

[168] S. Weiss, J. Eckel, M. Thorwart, and R. Egger, Itera-tive real-time path integral approach to nonequilibriumquantum transport, Phys. Rev. B 77 (2008), 195316.

[169] M. Thorwart, P. Reimann and P. Hanggi, Iterative al-gorithm versus analytic solutions of the parametricallydriven dissipative quantum harmonic oscillator, Phys.Rev. E 62 (2000) pp. 5808 - 5817.

[170] M. Thorwart, E. Paladino, and M. Grifoni, Dynamicsof the spin-boson model with a structured environment,

Chem. Phys. 296 (2004), pp. 333 - 344.[171] P. Nalbach, A. Ishizaki, G.R. Fleming, and M. Thor-

wart, Iterative path-integral algorithm versus cumulanttime-nonlocal master equation approach for dissipativebiomolecular exciton transport, New J. Phys. 13 (2011),063040.

[172] D.P.S. McCutcheon and A. Nazir, Coherent and inco-herent dynamics in excitonic energy transfer: Corre-lated fluctuations and off-resonance effects, Phys. Rev.B 83 (2011), 165101.

[173] D.P.S. McCutcheon and A. Nazir, Consistent treatmentof coherent and incoherent energy transfer dynamics us-ing a variational master equation, J. Chem. Phys. 135(2011), 114501.

[174] A. Kolli, A. Nazir, and A. Olaya-Castro, Electronicexcitation dynamics in multichromophoric systems de-scribed via a polaron-representation master equation, J.Chem. Phys. 135 (2011), 154112.

[175] G. Stock and M. Thoss, Mapping approach to the semi-classical description of nonadiabatic quantum dynamics,Phys. Rev. A 59 (1999), pp. 564 - 579.

[176] M. Thoss and G. Stock, Semiclassical description ofnonadiabatic quantum dynamics, Phys. Rev. Lett. 78(1997), pp. 578 - 581.

[177] E.R. Dunkel, S. Bonella, and D.F. Coker, Iterative lin-earized approach to nonadiabatic dynamics, J. Chem.Phys. 129 (2008), 114106.

[178] P. Huo and D.F. Coker, Iterative linearized density ma-trix propagation for modeling coherent excitation energytransfer in photosynthetic light harvesting, J. Chem.Phys. 133 (2010), 184108.

[179] A. Shabani, M. Mohseni, H. Rabitz, and S. Lloyd, Ef-ficient estimation of energy transfer efficiency in light-harvesting complexes, Phys. Rev. E 86 (2012), 011915.

[180] H.-P. Breuer and F. Petruccione, The Theory of OpenQuantum Systems, Oxford University Press, New York,2002.

[181] J. Roden, A. Eisfeld, W. Wolff, and W.T. Strunz, In-fluence of Complex Exciton-Phonon Coupling on Opti-cal Absorption and Energy Transfer of Quantum Aggre-gates, Phys. Rev. Lett. 103 (2009), 058301.

[182] G. Ritschel, J. Roden, W.T. Strunz, and A. Eis-feld, An efficient method to calculate excitation en-ergy transfer in light-harvesting systems: application tothe Fenna-Matthews-Olson complex, New J. Phys. 13(2011), 113034.

[183] L.A. Pachon and P. Brumer, Computational methodolo-gies and physical insights into electronic energy trans-fer in photosynthetic light-harvesting complexes, Phys.Chem. Chem. Phys. 14 (2012), pp. 10094 - 10108.