The systems and methods disclosed herein relate generally to the application of certain quantum coherent properties to the performance of various functions, such as providing a transition element for logical switching.
Moore's law famously predicts an exponential growth of computational power available through an ever-increasing number of transistors in a single electronic integrated circuit. However, as the 10 nm scale is approached in integrated circuit miniaturization, quantum tunneling effects start to limit the ability to increase transistor density in an electronic, semiconductor-based integrated circuit. Even before that scale is reached, excessive heat creation of ohmic wires virtually halted miniaturization of CPU units and forced manufacturers to produce multi-core processors instead. Photonics have been offered as one solution for overcoming the various limitations of electronic circuits. However, as computation is a complex multi-signal non-linear process, the realization of optical logic requires advanced non-linear optics currently not available. Thus, there is a need for alternative architectures for achieving nano-scale electronic circuits.
Some embodiments include wires, circuits, or circuit elements comprised of one or more chromophores. For example, a nano-scale wire may be constructed by a linear chain of chromophores supported by a substrate. In some embodiments, the chromophores are “tuned” to the critical edge between quantum order and quantum chaos. While not being bound by any particular theory, it is believed the chromophores at critical quantum chaos exhibit the unique property of long coherence times combined with quantum delocalization resulting in coherent transport of excitons. Thus, in some embodiments, an input signal at one chromophore (e.g., light stimulation) generates an exciton that can then coherently transport to adjacent chromophores in a nearly frictionless manner, allowing transport of energy and information with little heat generation. In some embodiments, such tuned chromophore systems provide coherent transport at room temperature. In some embodiments, the chromophores can have a configuration that places the system in a localization-delocalization threshold, and can facilitate keeping the system in the localization-delocalization threshold, thereby enabling coherent transport of excitons.
Some embodiments include an excitonic logic gate or ‘excitonic transistor’ where the timing and intensity of incident photons serve as a switching function that affects the transport of excitons through the gate in a manner similar to the voltage response of an electronic transistor. In some embodiments, such a gate includes two or more chromophores between which an incoming exciton would oscillate. The timing and intensity of incident photons modulates the probability of the gate to pass through the exciton.
Various embodiments relate to an information or energy conveyance structure that includes a chromophore assembly with a plurality of chromophores. In some embodiments, the plurality of chromophores can have a spatial configuration that paces the chromophore assembly within a pre-determined range of quantum order. The predetermined range of quantum order can include a critical transition point between quantum order and quantum chaos. In some embodiments, the plurality of chromophores can have a spatial configuration that paces the chromophore assembly within a pre-determined range of a metal-insulator transition. In some embodiments, the plurality of chromophores can have a spatial configuration that paces the chromophore assembly within a pre-determined range of a localization-delocalization transition.
Various embodiments relate to an apparatus for performing logical operations. The apparatus can include a chromophore assembly that includes a plurality of chromophores. The apparatus can include an exciton source configured to input an exciton into the chromophore assembly. The apparatus can include a chromophore modulator configured to modulate the probability that at least one chromophore in the chromophore assembly will transmit the exciton. In some embodiments, the plurality of chromophores can have a spatial configuration that paces the chromophore assembly within a pre-determined range of quantum order. The predetermined range of quantum order can include a critical transition point between quantum order and quantum chaos. In some embodiments, the plurality of chromophores can have a spatial configuration that paces the chromophore assembly within a pre-determined range of a metal-insulator transition. In some embodiments, the plurality of chromophores can have a spatial configuration that paces the chromophore assembly within a pre-determined range of a localization-delocalization transition.
Various embodiments relate to an apparatus for performing logical operations including a first module configured to apply external driving to a transmission element such that the transmission element's degree of quantum coherence exceeds a first threshold. The first module can drive the transmission element based on a first input that is configured to receive quantum information. The apparatus can include a second module that can be configured to maintain a state associated with the transmission element within a range of a transition point. The apparatus can include a third module configured to apply time dependent forces to the transmission element thereby reducing the transmission element's degree of quantum coherence below a threshold. The apparatus can include an output configured to transmit quantum information.
Various embodiments relate to an apparatus for performing logical operations. The apparatus can include an exciton source and an output. A first chromophore structure can be coupled between the exciton source and the output. The apparatus can include a first chromophore modulator, which can be configured to modulate the first chromophore structure between an open state and a closed state. A second chromophore structure can be coupled between the exciton source and the output. The apparatus can include a second chromophore modulator, which can be configured to modulate the second chromophore structure between an open state and a closed state.
In some embodiments, the apparatus can be configured such that an exciton is transferred from the exciton source to the output when either the first chromophore structure or the second chromophore structure is in the open state.
In some embodiments, the apparatus can be configured such that an exciton is transferred from the exciton source to the output when both the first chromophore structure and the second chromophore structure are in the open state, and such that the exciton is not transferred from the exciton source to the output when either the first chromophore structure or the second chromophore structure is in the closed state.
In some embodiments, the first chromophore modulator can be configured to modulate the first chromophore structure between the open state and the closed state by adjusting quantum coherence of the first chromophore structure.
Implementations disclosed herein provide systems, methods and apparatus for generating “excitonic transistors” and similar devices. Discovery of room temperature quantum coherence in the avian compass of birds, in the olfactory receptors and in light harvesting complexes in the last few years indicate that quantum effects might be ubiquitous in biological systems. While the quantum chemical understanding of certain details of light harvesting systems is almost complete, no organizing principle has been found which could explain why quantum coherence is maintained in these systems for much longer than the characteristic decoherence time imposed by their room temperature environment. Here we propose that at the critical edge of quantum chaos, coherence and transport can coexist for several orders of magnitude longer than in simple quantum systems. Quantum systems changing from integrable to quantum chaos pass through critical quantum chaos which is also a metal-insulator transition from Anderson localization to extended wave functions. By extending the semiclassical theory of decoherence from chaotic and integrable systems to the transition region we show that coherence decay changes from exponential to power law behavior and coherence time is amplified exponentially from its environmentally determined value. This result can be demonstrated on a ring of chromophores passing through the critical point where coherence in the critical point decays with the same non-trivial power law as in the FMO complex experiment. Our results also show that loss of coherence is not permanent in these systems and they can re-cohere via coherent external driving such as the arrival of photons in case of the light harvesting systems and can continuously hover in a “Poised Realm” between the coherent quantum and the incoherent classical worlds (see PCT Application No. PCT/US2011/044738 for discussion of this “Poised Realm”, which is incorporated herein by reference in its entirety). Some embodiments include using this new critical design principle from biology to build lossless quantum coherent energy and information processing devices operating at room temperature.
Quantum biology is dealing with open quantum systems closely coupled to their many degrees of freedom environment. The environment exerts time dependent forces on the system through the coupling. Some of these forces change very rapidly compared to the excitation frequencies of the system and look random from its point of view. This “heat bath” destroys quantum coherence and moves the system into a mixed state rapidly. The average effect of the random forces can be described as a non-unitary time evolution of the system's density matrix.
At room temperature the phonon environment has characteristic energy ET=kBT≈4·10−21J and frequency νr=kBT/h≈6 THz in the infrared spectrum. The typical phonon thermal wavelength is determined by the speed of sound of the material the system is embedded in. Its typical value for water and non-specific proteins is c=1500 m/s yielding λT=c/ν˜2.5 Å. Molecules within a thermal wavelength distance can feel the same environment and some of their states can be almost decoherence free and can preserve coherence for a relatively long time. In light harvesting systems the light absorbing and emitting chromophores are embedded in a protein scaffold which can suppress thermal fluctuations more effectively. The thermal wavelength can be increased to 10-13 Å by raising the effective speed of sound with a factor of 3-5 to 7-8000 m/s. Protective environments cannot extend the size of coherent patches further and a different mechanism is needed to extend it further, which we propose next.
The speed of environmental decoherence can be characterized by the decay rate of the off diagonal (n≠m) elements of the reduced density matrix of the system ρnm˜e−t/τc, where τc is the coherence time. Purity P=Tr[ρ2]=Σmn|ρnm|2 has been shown to be a good overall measure. It is P=1 when the system is in a pure state and decreases monotonically as the system decoheres into a mixed state. P(t)˜1/N+const·e−t/τ
Zero Lyapunov exponent and entropy production can also emerge in systems at the border of the onset of global chaos in the classical counterpart of the system. Suppose we have a parameter ε of the mechanical system which characterizes its transition from regular dynamics to chaos. H=HR+εHC, where HR is the Hamiltonian of a fully integrable system and HC is fully chaotic. Classically HR is a solvable system and it can be described by actionangle variables. It does only simple oscillations in the angle variables while the action variables do not change and remain conserved restricting the dynamics for the surface of a torus in the phase-space. When ε≠0 but small the system is not integrable classically and the Kolmogorov-Arnold-Moser (KAM) theory describes the system. The chaotic perturbation breaks up some of the regular tori in the phase-space and chaotic diffusion emerges localized between unbroken, so called KAM tori. Chaotic regions are localized in small patches in the phase-space surrounded by regular boundaries represented by the KAM tori. At a given critical εc even the last KAM tori separating the system gets broken and the chaotic patches merge into a single extended chaotic sea. In the transition region ε≈εc the Lyapunov exponent shows a second order phase transition with power law scaling λ0(ε)˜(ε−εc)β slightly above ε>εc with some exponent β>0. Above the transition ε>εc the system is chaotic characterized by a positive largest Lyapunov exponent λ0>0.
On the quantum mechanical level we can follow the transition in the statistical distribution of the energy levels. The Hamilton operator of the regular system HR is a separable with diagonal 4 matrix elements. The consecutive energy levels of the regular system look random and follow a Poisson process. The nearest neighbor level spacing distribution is then exponential p(s)=exp(−s), where sn=(En+1−En)/Δ(En) is the level spacing measured in the units of local mean level spacing Δ(E) at energy E. The Hamiltonian operator HC corresponding to the fully chaotic system is best approximated by a random matrix. The energy level statistics of Hc can be described by Random Matrix Theory (RMT) and the level spacings follow the Wigner level spacing distribution p(s)=(π/2) exp(−πs2/4) in systems with time reversal symmetry. As the parameter c is increased from zero the level spacing statistics changes from a Poissonian to a Wigner distribution. Critical quantum chaos sets up at the critical value εc in between. Below the critical point p(0) is finite, at the critical point and above the spacing distribution starts linearly p(s)=As for s→0, a characteristic feature of chaotic systems with strongly overlapping eigenfunctions. The tail of the distribution remains exponential below the critical point, exp(−Bs) for s→∞, which is a characteristics of regular systems whose eigenfunctions do not overlap for larger energy separations. It turns to Gaussian, exp(−Cs2), then above the critical point. At the critical point the level statistics is semi-Poisson, p(s)=4s exp(−2s), which starts linearly and decays exponentially combining the two main aspects of the level statistics of regular and chaotic systems.
The transition described here is more general than just the transition from regular to quantum chaotic behavior. It is also a transition from the localized states of the regular system to the extended states of the chaotic system. The two are separated by the metal-insulator transition (MIT) point between quantum mechanical Anderson localization and globally delocalized metallic phase. The transition point can be identified with the emergence of the semi-Poissonian level statistics. In the transition point the wave functions are neither fully localized nor extended and have an intriguing multi-fractal spatial character. The fractal structure allows them to develop a hairy localized structure but also an extended structure with long range overlap correlations.
Merging the pieces of classical, semiclassical and quantum aspects a new picture emerges. Systems well below the critical point have non-chaotic dynamics with zero generalized Lyapunov exponents and quantum localization lengths extending only for few states. Decoherence in these systems is slow and purity follows a power law decay P(t)≈1/tα with some exponent a making possible the presence of anomalously long living coherent dynamics in the system. But coherently evolving states remain localized and long distance quantum coherent transport is not possible. Systems well above the critical point have chaotic dynamics with positive Lyapunov exponents and delocalized states extending for the entire system. Coherence dies out exponentially fast. Near the critical point exponential decay of coherence crosses over to long living power law behavior and localized states become delocalized. In finite systems there is always a narrow region around criticality, where long living coherence and sufficiently extended states can exist at the same time.
This result can be demonstrated on a simplified model of chromophores in light harvesting complexes. It is very likely that biological systems use this mechanism to tune their parameters near the critical point to maintain rich quantum transport properties. The excitonic states are described in the single excitation approximation by the Hamiltonian Hij=ΣiEi|ii|++ΣijVij|i
j| where |i
indexes the excitonic states with site energies Ei and dipole interaction strengths Vij. For simplicity we take a simple ring of L chromophores coupled by constant Vnm=1 for neighboring sites n and m and zero otherwise and take quasi random on site energies given by En=W cos(2πσn), where the irrational number σ=(√5−1)/2 is the golden mean. This Hamiltonian is known as the one dimensional Harper model. At Wc=2 the infinite system L→∞ goes through a MIT with delocalized states below and localized states above criticality. At the critical point it has been shown to have semi-Poisson level statistics. The system is coupled to the phononic environment via the Hamiltonian ΣiF(xi, t)|1
i|, where F(x, t) is the randomly fluctuating phonon field including the chromophore site energy coupling constant. The reduced density matrix of the chromophore system can be described in Markovian approximation by the Lindblad equation
where Cnm=F(xn,t)F(xm, t)
is the correlation function of the environmental coupling. We assume that the correlation function depends only on the periodic distance of the chromophores in a simple way Cnm=D(L/π)2 cos2(π(n−m)/L) and is quadratic for small distances. Next we show results for L=25 (in dimensionless units
=1), which is a realistic number of chromophores in experimentally investigated systems. In
At criticality not only purity changes from exponential to power law decay but so does the population of the chromophores. In pn(t)
n˜t−1/4 at the critical point of MIT. The return probability p(t)=
|ψn(t)|2
is the probability of return assuming that the wave function was localized on the site initially ψn(0)=1. In the decoherence free case it coincides with the density matrix element ρnn(t) assuming ρnn(0)=1, which is shown in our model and for the FMO complex. It seems likely that the FMO complex follows the universal scaling of critical MIT indicating that the Hamiltonian of the FMO complex is tuned to critical quantum chaos in order to realize optimal coherent transport, what we show elsewhere.
The findings support a new approach to quantum biological systems. They are not just under the influence of environmental decoherence due to random noise but also driven by the coherent waves of the incoming photons. The photons are absorbed by one of the chromophores which can be interpreted as a measurement process selecting one of the chromophores randomly. Then the system is set into an initial state which is concentrated on the selected chromophore. The purity of the system becomes P=1 as this is a pure state and the partially decoherent evolution starts again decreasing the purity in time. The system can hover in the “Poised Realm” between clean quantum and incoherent classical worlds (see PCT Application No. PCT/US2011/044738, which is incorporated herein by reference in its entirety). By tuning the timings of re-coherence events and the coherence time during decoherence via tuning the system on the chaos-regularity axis can be kept in high level of purity. This makes it possible to create new quantum devices working at room temperature capable of nearly frictionless quantum transport of energy and information.
In some instances, quantum dissipation can play an important or essential role in the quantum search performed in the FMO complex. While is has been conjectured that light harvesting complexes such as the Fenna-Matthews-Olson (FMO) complex in green sulfer bacteria may perform an efficient quantum search similar to the quantum Grover's algorithm, the analogy has not been clearly established. The quantum search performed in the FMO complex is fundamentally different from Govenor's algorithm. It is something new not considered in quantum computation before. In the FMO complex not just the optimal level of phase breaking is present to avoid both quantum localization and Zeno trapping but its evolutionary design can harness quantum dissipation as well to speed the process even further up. With detailed quantum calculations taking into account both phase breaking and quantum dissipation we show that the design of the FMO complex has been evolutionarily optimized and works faster than pure quantum or classical-stochastic algorithms. Some embodiments utilize this new computational concept poised between the quantum and classical realms. The new computational devices can also be realized on different material basis, opening new magnitude scales for miniaturization and speed. In passing, we also derive a new equation generalizing the Caldeira-Leggett equation of quantum dissipation for arbitrary system Hamiltonians and system-bath couplings.
Some biological systems can benefit from quantum effects even at room temperature. It has been shown experimentally that quantum coherence can stay alive for an anomalously long time in light harvesting complexes. In these systems excitons initiated by the incoming photons should travel really fast throughout a chain of chromophores in order to reach the reaction center where they are converted to chemical energy. Excitons decay within 1 nanosecond and dissipate their energy back to the environment if they cannot find the photosynthetic reaction center via random hopping on the chromophores within that characteristic time. With classical diffusion via thermal hopping that time is easily consumed and evolution should have found more optimal ways to reach that goal. Quantum mechanics is very helpful in this respect as it allows the system to explore many alternative paths in parallel and can discover the optimal one faster than a classical random search would do. However, quantum mechanics has adverse effects too. Anderson localization can prevent excitons to travel large distances from their origins. Coupling the system to the environment breaks phase coherence and can destroy the negative effects of quantum localization. Too much phase breaking however slows down the propagation again via the Zeno effect. At the right amount of phase breaking environmental decoherence and quantum evolution collaborate to achieve optimal performance and efficiency. The Environment Assisted Quantum Transport (ENAQT) theory accounts for the interplay of these two effects and can explain the existence of a transport efficiency optimum at room temperature relative to both pure quantum or pure classical transport. ENAQT explains the quick quantum exploration of the search space at optimal phase breaking. Once the exciton can reach nearly ergodically the chromophore sites random trapping delivers of the exciton to the reaction center.
While ENAQT assures the fast spreading of probability over the light harvesting complex, it does not guide the exciton to the reaction center. The reason for this is that quantum mechanics and phase breaking leads to a uniform probability distribution over the state space. The reduced density matrix of a system with Hamiltonian H is described by the Lindbad equation
where the operators Vj describe the coupling of the system and the environment. In light harvesting systems the Hamiltonian Hnm is a discrete, where the chromophore sites are indexed by n=1, . . . , N. In case the chromophores are coupled to the environment independently the generators are simply diagonal Vj=√{square root over (γφ)}·|jj|, where γφ is the rate of phase breaking. The Lindblad equation keeps the density matrix normalized during the evolution Tr{ρ}=1 and its diagonal elements ρnm stay positive and give the probability of finding the exciton on site n. At the optimal level of phase breaking the system relaxes quickly to the uniform probability distribution ρnm=1/N. Trapping to the reaction center is described by the imaginary Hamiltonian −ih-κ|
r|, where r is the site of the reaction center and κ is the trapping rate. Assuming rapid relaxation to the uniform distribution the bulk of the time an exciton needs to get trapped by the reaction center is determined by the fraction of time it spends on the chromophore of the reaction center. The reaction center is able to catch an exciton sitting on it in average time 1/κ and the exciton spends ρrr fraction of its time on the chromophore. The average transport time is then the product
τ
≈1/(ρrrκ)=N/κ. One of the best studied light harvesting systems is the FMO complex which consists of N=7 chromophores. We use this example in some cases herein. It has been shown that ENAQT is optimal in this system at phase breaking rates of γφ=300 cm−1 corresponding to room temperature. At trapping rate 1 ps−1 the exciton needs about 7 ps to reach the reaction center, which is consistent with this estimate.
Since at optimal phase breaking the transport time depends only on the number of sites and on the trapping rate the concrete form of the FMO Hamiltonian plays no role as long as the relaxation to the uniform distribution is sufficiently fast. Accordingly, Hamiltonians with extended wave functions should be slightly more efficient than localized systems since the exciton is not trapped and the relaxation to the uniform distribution is somewhat faster. We demonstrate this in case of the FMO complex where the Hamiltonian Hnm has been obtained from spectroscopy. The diagonal part of the Hamiltonian consists of the site energies of the chromophores. The off diagonal hopping terms describe the transition between sites. We can modify the localization properties of this Hamiltonian by rescaling the diagonal elements relative to the off diagonal elements H′nm=Hmm+(λ−1)δnmHnm, where λ, is the tuning parameter. For λ=1 we recover the original Hamiltonian. For λ>I the diagonal elements become larger and the system becomes completely localized for λ→∞, while for 0≦γ<1 the system becomes more extended.
One way to study the relaxation to the correct thermal equilibrium is to use the Redfield equations describing the interaction of the system and the environmental bath. The Redfield equation can be cast in a form similar to the Lindblad equation
where the operators can be written in energy representation as [Vj+]ab=[Vj]ab/(1+eβ(E
For high temperatures we can expand this equation for small β. The first two terms in the expansion arc basis independent
while the third term in the expansion is zero in general. The first term is the Lindblad equation for self-adjoint operators Vj. The second term describes quantum dissipation, which is missing from the Lindblad equation. Caldeira and Leggett (CL) showed that the reduced density matrix of open quantum systems coupled to a high temperature bath experience both phase breaking and quantum dissipation and satisfy the equation
Our new equation (4) gives back the CL equation as a special case for the Hamiltonian
with coupling V=√{square root over (γφ)}x and it is valid for a much larger class of Hamiltonians and operators V. In particular for discrete Hamiltonians Hnm describing the exciton dynamics in light harvesting complexes and for environmental couplings Vj=√{square root over (γφ)}·|jj| it takes the form
The most important feature of this equation is that the quantum dissipative term cannot be chosen arbitrarily in models of exciton dynamics. The Hamiltonian and the generators Vj determine both phase breaking and dissipation uniquely. Also the order of magnitude the dissipative term relative to the phase breaking term is determined by the ratio of the size of the typical Hamilton matrix element and the temperature. In light harvesting systems these are comparable and quantum dissipation cannot be neglected.
Quantum dissipation speeds up the transport process in light harvesting complexes. If the site energies at the reaction center are lower than in the other parts of the complex the equilibrium density is higher and the exciton spends longer time on the chromophore related to the reaction center and is trapped with higher probability. The average time is again τ
=1/(κρrr) but now the probability is given by the Boltzmann factor ρrr=
r|e−βH|
/Z . In case of the FMO complex this probability is about 40% and the transport time would drop to a mere 2.5 picoseconds in this approximation at optimal phase breaking. Our detailed calculation using the Redfield operators outlined in the supplementary material yields about 3.5 picoseconds which is very close to this estimate and less than half than it would be without quantum dissipation. We can now ask in what sense is this result optimal? Could we achieve a better result by choosing as deep site energy as possible so that ρrr≈1 can be achieved? We show next that this absolute optimum cannot be attained and the real FMO operates with the best transport time possible physically and evolutionarily.
At the optimal phase breaking of ENAQT quantum dissipation introduces a trade-off between fast relaxation to the equilibrium distribution and the shape of the equilibrium distribution. The equilibrium density matrix can be expressed in terms of the energy eigenstates ψn(k) as
If the system is completely delocalized the wave functions are extended |ψn(k)|2≈1/N and the diagonal elements of the density matrix become uniform ρnm≈1/N independently of the energy levels Ek of the system. In this case the relaxation to the equilibrium is fast since the extended wave functions overlap strongly with the exciton starting on one of the chromophores, but the exciton spends time on each chromophore nearly equally. If the system is strongly localized the wave functions are concentrated on single sites |ψn(k)|≈δnk and ρnm≈e−βE
We think that this picture is quite general. If we consider larger transport systems the optimum would again lie somewhere midway between the extended and localized cases. Since the localization-delocalization transition is getting sharper with increasing system size these systems can be found (e.g., may only be found) at parameters near the metal-insulator threshold. To demonstrate this in
In the light harvesting case, the task of the system is to transport the exciton the fastest possible way to the reaction center whose position is known. In a computational task we usually would like to find the minimum of some complex function ƒn. For the simplicity let this function have only discrete values from 0 to K. If we are able to map the values of this function to the electrostatic site energies of the chromophores Hnm=εQfn and we deploy reaction centers near to them trapping the excitons with some rate κ and can access the current at each reaction center it will be proportional with the probability to find the exciton on the chromophore jn˜κρnm. Since the excitons will explore the Boltzmann distribution the currents will reflect that jn=κn|e−βH|n
/Z . There are three conditions which should be valid simultaneously: 1) The system should operate at the optimal phase breaking which then should be in the order of magnitude of the energy steps γφ˜O(ε0). 2) In the worst case scenario the minimum current is elevated with a factor eβε0 relative to the second smallest minimum. To be able to detect this, the energies should be of the order of the thermal energy ε0˜O(kBT). 3) The hopping terms Hnm between the chromophores should be optimal to keep the system at the border of the localization-delocalization transition. The first two conditions can be easily met since the phase breaking is usually of the same order as the thermal energy γφ˜kBT. The third condition can be realized by placing the chromophores interacting via the dipole interaction to an optimal distance from each other randomly so that the quasi random Hnm matrix elements keep the system at the localization-delocalization threshold. Conversely, given a random arrangement of Hnm-s the parameter ε0 can be tuned so that the system gets to the localization-delocalization threshold. In some cases, the localization length can be calculated and the distance between the chromophores can be adjusted until a suitable localization length is provided (e.g., at or near the localization-delocalization threshold). For example, the distance between the chromophores can provide a localization length that is about halfway between fully localized and fully extended.
This quantum-classical optimization method discovered by evolution seems to be superior to the optimization methods developed so far. Classical stochastic optimization techniques can be trapped in local minima for long times and careful annealing techniques are required to reach the correct minimum. Even then sites are discovered in a classical sequential manner and it takes the process long times to find the minimum. Quantum mechanics is more advantageous as it is able to explore the sites in parallel, but the discovery process is hampered by Anderson localization especially near local minima. An optimal amount of phase breaking can destroy the interferences causing this and can ensure the ergodic exploration of the states while quantum dissipation takes all the advantages of the classical stochastic optimization and establishes the Boltzmann distribution which elevates the proper minimum. The physical speed of the process is determined by the inverse trapping rate 1/κ which is in the order of picoseconds.
Current computers operate with about 4 GHz processors, where the cycle time of logical operations is 250 picoseconds. Computers based on artificial light harvesting complexes could have units with 100-1000 times larger efficiency at room temperature. But, it is also possible to realize such systems on excitons of organic molecules or on Hamiltonians arising in nuclear matter, which would provide a virtually endless source of improvement both in time and miniaturization below the atomic scale.
In an alternative embodiment of
It will be appreciated that where the various structures depicted in
The nanocircuits illustrated in
The chromophore structures may be harvested from the biological entity and relocated to a controlled environment, such as a microfluidic system, conducive to the constructions of integrated logical circuits. The controlled environment may comprise a sheet having predetermined attachment points configured to attract the chromophore structures. Spraying the harvested chromophore structures upon the sheet would then result in the chromophores being affixed to the desired location, such as locations in communication with electrical contacts as discussed in relation to
In some embodiments, the chromophore structures used as described herein exhibit critical quantum chaos. Experimentally, chromophore structures existing in this state can be identified by measuring the decay of coherence in the system. In a non-critical system the coherence decay is exponential in time. Critical quantum chaos is signaled by a slow, typically power law decay of coherence. State of the art coherence decay measurements are based on various echo measurements depending on the system studied. This includes spin echo, neuton spin echo, and photon echo.
In some embodiments, whether a candidate chromophore structure exhibits critical quantum chaos can be determined by analyzing the energy level spacing distribution of its quantum degrees of freedom (e.g., by using spectroscopic techniques). In a pure ordered regime, the energy level spacing distribution has the form:
p(s)=exp(−s)
where s is the energy level spacing and p(s) is the energy level spacing distribution. In a purely chaotic system, the distribution has the form:
At critical quantum chaos, the energy level spacing distribution has the form:
p(s)=4s exp(−2s).
The degree of closeness of the system to critical quantum chaos may determined by calculating the value:
where pp(s)=exp(−s) and
p(s) is the experimentally determined energy level spacing distribution. The theoretical critical point is at x=0.475. In various embodiments, a chromophore assembly is provided with an x value between about 0.4 and about 0.6, between about 0.45 and about 0.55, between about 0.45 and about 0.5, and at about 0.475.
Further details regarding designing or identifying chromophore structures that exhibit critical quantum chaos may be found in PCT Application No. PCT/US2011/044738 (published as WO 2012/047356), which is incorporated herein by reference in its entirety.
Accordingly, in some embodiments, chromophore structures for use in the systems described herein are selected by analyzing rate of coherence decay and/or energy level spacing distribution of a selection of candidate chromophore structures. Those chromophore structures meeting the criteria described above are then used in the construction of a nanocircuit.
Those having skill in the art will further appreciate that the various illustrative logical blocks, modules, circuits, and process steps described in connection with the implementations disclosed herein may be implemented as electronic hardware, computer software, or combinations of both. To clearly illustrate this interchangeability of hardware and software, various illustrative components, blocks, modules, circuits, and steps have been described above generally in terms of their functionality. Whether such functionality is implemented as hardware or software depends upon the particular application and design constraints imposed on the overall system. Skilled artisans may implement the described functionality in varying ways for each particular application, but such implementation decisions should not be interpreted as causing a departure from the scope of the present invention. One skilled in the art will recognize that a portion, or a part, may comprise something less than, or equal to, a whole. For example, a portion of a collection of pixels may refer to a sub-collection of those pixels.
Headings are included herein for reference and to aid in locating various sections. These headings are not intended to limit the scope of the concepts described with respect thereto. Such concepts may have applicability throughout the entire specification.
The previous description of the disclosed implementations is provided to enable any person skilled in the art to make or use the present invention. Various modifications to these implementations will be readily apparent to those skilled in the art, and the generic principles defined herein may be applied to other implementations without departing from the spirit or scope of the invention. Thus, the present invention is not intended to be limited to the implementations shown herein but is to be accorded the widest scope consistent with the principles and novel features disclosed herein.
This application claims the benefit under 35 U.S.C. §119(e) of U.S. Provisional Patent Application No. 61/590,745, filed Jan. 25, 2012, and titled CHROMOPHORE BASED NANO-CIRCUITS, U.S. Provisional Patent Application No. 61/613,936, filed Mar. 21, 2012, and titled CHROMOPHORE BASED NANOCIRCUITS, and U.S. Provisional Patent Application No. 61/714,610, filed Oct. 16, 2012, and titled CHROMOPHORE BASED NANOCIRCUITS, each of which is hereby incorporated by reference in its entirety and made a part of this specification for all that it discloses.
Number | Date | Country | |
---|---|---|---|
61590745 | Jan 2012 | US | |
61613936 | Mar 2012 | US | |
61714610 | Oct 2012 | US |