Anda di halaman 1dari 18

Emission of Biophotons and Neural Activity of the Brain

1,2,3
Majid Rahnama, 4Istvan Bokkon, 5,6Jack Tuszynski, 7,8Michal Cifra,
1
Peyman Sardar, 1,9,10Vahid Salari*
1
Department of Physics, Shahid Bahonar University of Kerman, Kerman, Iran
2
Kerman Neuroscience Research Center (KNRC), Kerman, Iran
3
Afzal Research Institute, Kerman, Iran
4
Doctoral School of Pharmaceutical and Pharmacological Sciences, Semmelweis University, Hungary
5
Department of Experimental Oncology, Cross Cancer Institute, 11560 University Avenue, Edmonton, AB T6G 1Z2, Canada
6
Department of Physics, University of Alberta, Edmonton, T6G 2J1, Canada
7
Institute of Photonics and Electronics, Academy of Sciences of the Czech Republic, Prague, Czech Republic
8
Department of Electromagnetic Field, Faculty of Electrical Engineering, Czech Technical University in Prague, Prague,
Czech Republic
9
Institut de Mineralogie et de Physique des Milieux Condenses, UPMC-Paris 6, CNRS UMR7590, France
10
BPC Signal, 15 rue Vauquelin, 75005 Paris, France

* Corrsponding Author, Email Address: vahidsalari@impmc.upmc.fr

Abstract

In this paper, we argue that, in addition to electrical and chemical signals propagating in the neurons of the
brain, signal propagation takes place in the form of biophoton production. This statement is supported by recent
experimental confirmation of photon guiding properties of a single neuron. We have reviewed the mechanism of
biophoton production in the neurons and investigated the interaction of biophotons with biomolecules from a
quantum mechanical point of view. It appears that the role of biophotons in the neurons of the brain is very
important. A significant relationship between the fluctuation function of biomolecules (due to the emission and
absorption of biophotons) and alpha-EEG diagrams is elaborated on in this paper.

Keywords: mitochondrial biophoton, microtubule (MT), coherence, fluctuation function

1. Introduction

Biophotons, or ultraweak photon emissions of biological systems, are weak electromagnetic waves in
the optical range of the spectrum (i.e. light). All living cells of plants, animals and humans emit
biophotons which cannot be seen by the naked eye but can be measured by special equipment 1. Living
cells spontaneously emit ultraweak light during the process of metabolic reactions associated with
their physiological states. Neural cells also continuously emit biophotons during their natural
metabolism. There is a direct correlation between the intensity of biophotons and neural metabolic
activity in hippocampal slices of rat brain (Isojima et al. 1995).
In in vivo experiments, the spontaneous ultraweak biophoton emission from a rat brain was
shown to correlate with cerebral energy metabolism, EEG activity, cerebral blood flow and oxidative
stress ( Kobayashi et al. 1999). Van Wijk et al.(2008) performed an experiment using the technique of

1
Updated from: http://www.transpersonal.de/mbischof/englisch/webbookeng.htm, 23 November 2010.

1
“Ultra-weak Photon Emission” (UPE) to investigate the correlations between the biophoton emission
and the electrical activity of the brain. Significant correlations were demonstrated between the
fluctuations in photon emission and fluctuations in the strength of alpha wave production. Some
unpublished observations suggest that the state of the biophoton field of a person may be connected to
the state of the brain as measured by the EEG (e.g., degree of synchronization and coherence) (
Bischof 2005). Advanced states of deep relaxation or certain meditative states characterized by a high
degree of coherence in the EEG measurements may well be accompanied by a high degree of
coherence in the biophoton field. It is not unreasonable to hypothesize that even the visual field is a
property of the biophoton field itself (Bischof 2005). Bischof (2005) cautioned that these are merely
conjectures because measurements correlating the coherence of the biophoton field and the EEG
readings have not been made yet. In general, there is a neural activity-dependent biophoton emission
within neuronal cells, but evidence for the correlation between the electrical activity of the neurons
and the biophoton emissions is unfortunately quite poor. In this paper, we have investigated
theoretically the interaction of biomolecules with biophotons taking place within the neurons of the
brain. We have adopted a quantum mechanical formalism in an attempt to investigate possible
connections between the EEG and the biophoton production.

2. Biophoton production mechanism inside the neuron

During natural metabolic processes taking place in various kinds of living organism, permanent and
spontaneous ultraweak biophoton emission has been observed without any external excitation
(Isojima et al. 1995, Kobayashi et al. 1999, Van Wijk et al. 2008, Bischof 2005, Adamo et al. 1989,
Tilbury and Cluickenden 1988, Quickenden and Que Hee 1974, Kobayashi et al. 1999, Van Wijk and
Schamhart 1988, Mansfield 2005, Scott et al. 1991, Takeda et al. 1998, Yoon et al. 2005, Grass and
Kasper 2008, Imaizumi et al. 1984, Popp et al. 1984). Ultraweak photon emission from unicellular
and multicellular organisms is referred to using various terms such as dark luminescence, low
intensity chemiluminescence, ultraweak electromagnetic light, bioluminescence, ultraweak photons,
biophotons etc. The emergence of biophotons is due to the bioluminescent radical and nonradical
reactions of Reactive Oxygen Species (ROS) and Reactive Nitrogen Species (RNS), and involves
simple cessation of excited states. As examples we can cite the mitochondrial respiration chain and
peroxisomal reactions, non-enzymatic and enzymatic lipid peroxidation, oxidation of catecholamines,
oxidation of tyrosine and tryptophan residues in proteins, etc. (Watts et al. 1995, Kruk et al. 1989,
Nakano 1989). The main source of biophotons derives from the oxidative metabolism of mitochondria
(Tuszynski and Dixon 2001).

Ultraweak biophoton emission from neural tissue depends on the neuronal membrane
depolarization and Ca2+ entry into the cells (Kataoka et al. 2001). This biophoton emission can be
facilitated by the membrane depolarization of neurons by a high concentration of K+ and can be
attenuated by application of tetrodotoxin or removal of extracellular Ca2+ (Kataoka et al. 2001). The
photon emission intensity from the brain slices depends on temperature and oxygen concentration.
Therefore, we can conclude that biophoton emission within neurons is directly correlated with
biochemical processes.

However, the term ultraweak luminescence (ultraweak biophoton emission) can be


misleading, since it may suggest that biophotons are not important for cellular processes. We should
stress that the real biophoton intensity within cells and neurons can be considerably higher (Bokkon et
al., 2010) than one would expect from the measurement of ultraweak bioluminescence, which is
generally carried out macroscopically several centimeters away from the tissue or cell culture (Thar
and Kühl 2004). Specifically, the most significant fraction of natural biophoton intensity cannot be
measurable because it is absorbed during cellular processes. Moreover, it is a well-known fact that
photons are strongly scattered and absorbed in cellular systems. Estimates indicate that for a measured
intensity of the ultraweak bioluminescence (biophotons), the corresponding intensity of the light field
within the organism can be up to two orders of magnitude higher (Chwirot 1992, Slawinski 1988).

2
Biophoton communication (optical communication) can be performed by natural
photosensitive biomolecules of cells and neurons. Biophotons can be absorbed in the visible range by
natural chromophores such as porphyrin rings, flavinic, pyridinic rings, lipid chromophores, aromatic
amino acids, etc (Thar and Kühl 2004, Kato et al. 1981, Karu 1999, Mazhul' and Shcherbin 1999).
Mitochondrial electron transport chains contain several chromophores, among which cytochrome
oxidase enzyme are most prominent (Kato et al. 1981, Karu 1999). Photosensitive biomolecules can
transfer the absorbed biophoton energy to nearby biomolecules by resonant energy transfer, which can
induce conformational changes and trigger/regulate complex signal processes in cells. According to
Cilento 1988, biochemical reactions take place in such a way, that a biophoton is borrowed from the
surrounding electromagnetic bath, and then it excites the transition state complex, and finally returns
to an equilibrium state within the surroundings of the complex.

Recently, Sun et al. (2010) experimentally demonstrated that neurons can conduct photon
signals. Sun and colleagues suggested that bioelectronic and biophotonic processes are not
independent biological events in the nervous system, and their synergistic action may play a
significant role in neural signal transductions. Moreover, Wang et al.(2010) recently presented the
first experimental proof of the existence of spontaneous biophoton emission and visible light induced
delayed ultraweak photon emission. In their experiments they used in vitro freshly isolated rat’s whole
eye, lens, vitreous humor and retina. As a consequence of their findings they proposed that the
photochemical source of retinal discrete noise, as well as retinal phosphenes, may originate from
natural bioluminescent photons within the eyes (Bokkon, 2003, 2005, 2008, 2009).

Numerous findings have provided evidence of fundamental roles of ROS and RNS in cellular
processes (Dröge 2002, Valko et al. 2007). Under physiological conditions, ROS and RNS can control
gene expression, apoptosis, cell growth, cell adhesion, chemotaxis, protein-protein interactions and
enzymatic functions, Ca2+ and redox homeostasis, as well as numerous other processes of cells (Dröge
2002, Valko et al. 2007, Touyz 2005, Gordeeva et al. 2003, Chiarugi 2003, 2008, Hoidal 2001). There
are experimental indications that ROS and RNS are also necessary for synaptic processes and normal
brain functions. Free radicals and their derivatives act as signaling molecules in cerebral circulation
and are necessary for molecular signaling processes in the brain such as synaptic plasticity,
neurotransmitter release, hippocampal long-term potentiation, memory formation, etc. (Volterra et al.
1994, Kishida and Klann 2007, Knapp and Klann 2002, Tejada-Simon et al. 2005, Thiels et al. 2000,
Thiels and Klann 2002).

Because the generation of ROS and RNS is not a random process, but rather a co-ordinated
mechanism used in signaling pathways under various physiological conditions, biophoton emission
may not be a byproduct of biochemical processes but it can be linked to precise signaling pathways of
ROS and RNS. Consequently, during natural oxidative metabolism, regulated generation of ROS and
RNS can also generate regulated biophotons within cells and neurons. This means that regulated
electrical (redox) signals (spike-related electrical signals along classical axonal-dendritic pathways)
of neurons can be converted into biophoton signals by various bioluminescent reactions. As a result,
information appears not only as an electrical signal but also as a regulated biophoton (optical) signal
in the brain.

3. Interaction of Biophotons with Microtubules (MTs)

In this section, we theoretically investigate the mechanism involved in the interactions of biophotons
with MTs.

Microtubular structures have been implicated as playing an important role in the signal and
information processing taking place in the human (and possibly animal) brain (Jibu et al. 1994, 1996,
Hameroff and Penrose 1996, Mavromatos et al. 1998, 2002). Vertebrate cells and neurons show
typically filamentous mitochondria associated with the microtubules (MTs) of the cytoskeleton,
forming together a continuous network (mitochondrial reticulum) (Skulachev 2001). The rapid
movements of mitochondria are MT-based and the slower movements are actin-based (Tong 2007).
MT formation can be regulated by redox-dependent phosphorylations and Ca2+ signals. Since the

3
rapid movements of mitochondria are MT-based, mitochondrial trafficking can be organized by redox
and Ca2+-dependent MT regulation. Moreover, the refractive index of both mitochondria and MTs is
higher than the surrounding cytoplasm (Thar and Kühl 2004). Consequently, both the mitochondria
and the MTs could act as optical waveguides, i.e. electromagnetic radiation can propagate within
their networks (Thar and Kühl 2004, Faber et al. 2006, Jibu et al. 1994). Since mitochondria and MTs
could act as an optical waveguide and because filamentous mitochondria are associated with the
MTs, it is logical to predict that mitochondrial biophotons can propagate along the associated
mitochondria-MTs dynamic networks. Associated mitochondrial and MT networks may act as redox
and Ca2+ regulated organic quantum optical-like fiber systems in neurons. MTs are composed of
tubulin dimers. Tubulin dimer is an intrinsically fluorescent molecule mainly due to 8 tryptophan
residues it contains, as can be seen e.g. in the 1TUB structure from the Protein Data Bank (Nogales et
al. 1998). It is well known that the absorption (ca. 280 nm) and fluorescence (ca. 335 nm)
wavelength (and intensity) of tryptophan is dependent on the conformation of tubulin. Probing
absorption and fluorescence of tubulin is a standard method to determine the polymerization state of
MTs. This can be considered one of possible qualitative connections between the fluctuations of MT
growth and its corresponding biophoton absorption and emission characteristics. Additionally, there
exist other energy states (of both optical and vibrational nature (Jelínek and Pokorný 2001, Pokorný
et al. 1997, Deriu et al. 2010) which tubulin and the whole microtubule can support. These states can
be excited by energy supply provided by mitochondria (Cifra et al. 2010). Furthermore, microtubule
polymerization is sensitive to UV (Staxén et al. 1993) and blue light (Murata et al. 1997) and
mitochondria are known to be sources of biophotons corresponding to the same wavelengths
(Vladimirov and Proskurnina 2009, Hideg et al. 1991, Batyanov 1984), which makes an immediate
logical connection. Thus we see that the tubulin is a protein which can interact (absorb and emit) with
photons that can be also related to a redox mechanism used in the mitochondria.

Here, we consider tubulins as representative biomolecules. Next, we examine what happens when the
biophotons interact with the microtubular structures in the neurons of the brain.
Earlier, MTs have been considered as optical cavities (Mavromatos et al. 2002) with quantum
properties (Mavromatos and Nanopoulos 1998), capable of supporting only a single mode (Jibu et al.
1996) or perhaps a few widely spaced (in the frequency domain) modes. Our approach is based on a
fully quantum formalism of the Jaynes-Cummings model (Jaynes and Cummings 1963). We consider
the tubulin dimer to represent a two-state system with ground g , and excited e states, respectively.
This system interacts with a single-mode cavity field of the form
1
   2
Eˆ  eˆ  a  a sin(kx)
† (1)
  0V 
where ê is an arbitrary oriented polarization vector. The interaction Hamiltonian is given by (Gerry
and Knight 2005)
Hˆ ( I )  dˆg (aˆ  aˆ † ) (2)
  
where g  (  ) 2 sin(kx ) and d  dˆ.eˆ .The quantity  er is the dipole moment of tubulin, d  er . In
 0V
general, i.e. for an unspecified representation, the dipole moment is an operator, d̂ . Here, we
introduce the transition operators as ̂   e g , ˆ   g e  ˆ † and the fluctuation
operator ̂ 3  e e  g g (Gerry and Knight 2005). The interaction Hamiltonian is:
H ( I )   (ˆ   ˆ  )( aˆ  aˆ † ) (3)
where   dg . Assuming that tubulin has only two well-spaced quantum states g and e , the

Hamiltonian for tubulin may be written as Hˆ tubulin  1 ( E e  E g )ˆ 3  1 0ˆ 3 . The energy difference
2 2
between the excited state and the ground state of tubulin is equivalent to the production or absorption
of a biophoton. For the field inside the MT, after dropping the vacuum energy, we have Hˆ f  a † a .
We conclude that the photons can be transmitted via MTs to the brain, thus the biophotons should be

4
the same as incident photons. In this case, 0   . With this assumption in place, the total
Hamiltonian becomes:
1 (4)
Hˆ  Hˆ tubulin  Hˆ f  Hˆ Interaction   0 ˆ 3  a † a   (  a    a † )
2
Here, the state vector of tubulin is Tubulin  c g (t ) g  ce (t ) e ) and the state of the field

is Field   cn n . The total state is the tensor product  (t )  Tubulin  Field . Inserting the total
n0

state into the Schrödinger’s equation i d  (t )  Hˆ Interaction  (t ) , the result is readily obtained as
dt

|  (t )   

  c c cos( t n  1 )  ic c
 e n g n 1 sin(t n  1 | e  

|n
 
n  0    ice cn 1 sin( t n )  ic g cn cos(  t n ) | g }
 
The Wu-Austin Hamiltonian has been used in the past to describe the interaction of quantized
electromagnetic field with MTs to yield a coherent Froehlich’s state of the dipolar biological system.
Wu and Austin (Wu and Austin 1977, 1978, 1981) proposed a dynamical model containing a
biological system composed of electric dipoles with N modes connected to harmonic heat baths
representing a quantized electromagnetic source and the surrounding thermal-relaxation bath. For N B
relaxation-bath modes k of frequency  k and N I input electromagnetic modes l of frequency 1' , the
model Hamiltonian is
N NB N I N NI
H Wu  Austin    i ai ai    k bk bk    'l cl cl   ai cl    ai cl
i 1 k 1 l 1 i 1 l 1
N NB NB N N NB
  ai b j bk    aib j bk  ai a j bk    ai a j bk
i 1 j 1 k 1 i 1 j 1 k 1

where ai , bk and cl are creation operators for the system, heat bath, and quantized electromagnetic
field, respectively. The parameters present in the Wu-Austin Hamiltonian are the coupling
constants  ,  , and  . We believe that the system of neuronal MTs is a good candidate for being
properly described by the above Hamiltonian. MTs are composed of tubulin dimers which can be
considered as biological electric dipoles. Previously one of the concerns of considering coherent
states in the brain was due to the fact that the Bose-Einstein condensation happens only at low
enough temperatures believed to be lower than body temperature. Recently, Reimers et al. 2009 have
argued that a very fragile Froehlich coherent state may paradoxically only emerge at very high
temperatures and thus there is no the possibility for the existence of Froehlich coherent states in
biological systems, so quantum models based on the Froehlich coherence should be ruled out. It has
been shown that there are serious problems in the calculations made by Reimers et al (2009) and
consequently their conclusions appear to be flawed (Salari et al, 2010).

However, an important problem that remains when considering coherent states for MTs is the so-
called decoherence problem. The question is “how is it possible for MTs to be in a coherent state
while the environment surrounding them is relatively hot, wet and noisy?” Tegmark (2000) calculated
decoherence times for MTs based on the collisions of ions with microtubules leading to the
decoherence times on the order of:

D 2 mkT (5)
  10 13 s
Ngq 2
where D is the tubulin diameter, m the mass of the ion, k Boltzmann’s constant, T absolute
temperature, N the number of elementary charges in the microtubule interacting system, g  1 the
4 0
Coulomb constant and q the charge of an electron. Hagan et al. showed that Tegmark used wrong
assumptions for his investigation of MTs. Another main objection to the estimate in (5) is that

5
Tegmark’s formulation yields decoherence times that increase with temperature contrary to a well-
established physical intuition and the observed behavior of quantum coherent states (Salari et al,
2010). In view of these (and other) problems in Tegmark’s estimates, Hagan et al. 2002 asserted that
the values of quantities in Tegmark’s relation are not correct and thus the decoherence time should be
approximately 1010 times greater. According to Hagan et al., MTs in neurons of the brain can process
information quantum mechanically and they could avoid decoherence via several mechanisms over
sufficiently long times for quantum processing to occur. As a result, we conclude that coherent states
in MTs are still theoretically possible. Below, we explore the consequences of this conclusion.

We begin by representing the state of the MTs as a superposition of the coherent and incoherent
states, and the state of the biophotons as a field composed of n photons. In our approach, transitions
involving the coherent-incoherent process or, analogously, the preconscious-conscious process are
determined by a function called the fluctuation function, F (t ) .

Now, consider two states of a MT: the ground state (incoherent without energy pumping) 0 , and an
excited state (coherent with energy pumping) z . We reformulate our previous calculations for the
system of MTs and investigate its interaction with biophotons. First, prior to the interaction, MTs
may be in one of the two states: the ground state or the coherent state. We continue our calculations
for each of these states.
The state of MTs can be written in the form of a superposition of the ground state and coherent state.
MTs vibrate in different frequency modes before the energy pumping occurs. This state can be
written in the form n1 n2 n3 ....  n1 , n2 , n3 ,...  n j  and the energy of MTs is E    J ( n J  1 ) .
J
2
These states are orthogonal according to n1 , n 2 ,... n , n ,...    .... . All these states can be
'
1
'
2 ' '
n ,n n n 1 1 2 2

† nj
generated from the vacuum according to  n   ( aˆ ) 0 . j
j  j nj!
MTs vibrate in different modes in the absence of electromagnetic waves, but when electromagnetic
waves are pumped into this system, according to Froehlich’s theory they occupy one frequency mode
which has a higher energy. This higher energy state is a coherent state. Assuming that before energy
pumping occurs, MTs are in the ground state with n j  0 , the energy is E    J ( 1 ) and
J
2
n 
j  0 . This is a living system and we let the ground state of the MT be equivalent to the ket 0 .
Now, the state of MT can be written as the superposition of the ground state and coherent state,
| (0) biomolecules  c g | 0  ce | z (6)
2
z
where z is our representation for the coherent state. It is written as z  e 

zn
2
 n .
n 0 n!
We now investigate its interaction with N biophotons. The state of biophotons is considered as

|  ( 0)   cn | n (7)
biophotons
n0

and the total state is


| (0) | (0)  |  ( 0) (8)
biomolecules biophotons

After solving the Schrödinger’s equation we have the time-dependent wave function as

| (t )   

  c c cos( t n  1)  ic c
 e n g n 1 sin(t n  1 | z 
 
|n

n  0    icecn 1 sin( t n )  ic g cn cos(  t n ) | 0 }
  
It is clear that this state is an entangled state, because it cannot be rewritten as a tensor product.
In this case, the MT system is considered as a system of biomolecules (tubulins), and it first occupies
the ground state. It means c g  1 , and ce  0 .

6
After the substitution of these coefficients, the state becomes

   ic
| (t )   
 
n 1 sin(t n  1 | z  
(9)
| n

n 0   ic n cos( t n ) | 0 }
  
Written in another form the state is
| (t ) |  g (t ) | 0  | e (t ) | z (10)
where

|  e (t )  i n 0 cn sin(t n ) | n  1 (11)

|  g (t )  i  c n cos( t n ) | n (12)
n0

Now, we introduce the fluctuation operator  3 as


 3 | e e  | g g  z z |  0 0 | (13)
To determine when the MT is in the coherent state and when in the ground state, we use the
fluctuation function. We define the fluctuation function as
F (t )    3 
This function determines the rate of transitions between the coherent state and the ground state.

Eventually the fluctuation function for this state is


2
 z 2 

n0
2 2
F (T , z )   c n cos ( n T ) e  z
2

 

n0
2


2
 1   cn sin ( n T ) e 2  e
z


(14)
 
where T  t . We refer to the parameter T as “scaled time”.
Here, cn is the coefficient describing the biophoton field. Biophotons have also been considered
coherent elsewhere (Bischof 2005, Popp and Yan 2002, Bajpai 1998, 1999), but the claim of their
coherence still requires a concerte proof. Here we hypothetically accept the coherence property of
biophotons and investigate its consequences. Thus for biophotons we have
2


2
n
cn  e
n!
We know that normalization implies that  2  N , where N is the average number of biophotons.
n
Thus, c n 2  e  N N .
n!
Finally, the fluctuation function becomes
2

N Nn  2 2
z 
F (T , z )   e 1  cos ( nT )  sin (e
2
nT )
n0 n!  

The 3-D diagrams of the fluctuation function are plotted in terms of the scaled time and coordinate z,
for different numbers of biophotons N. According to earlier calculations (Bokkon et al. 2010), at least
100 biophotons can be produced within each human visual neuron per second during visual
perception. So, a few biophotons per second are probably simultaneously present within the 1-10
micrometer length of a MT. Here, we have considered different values for the numbers of biophotons
around a MT such as N=2, 5, 7 and 10 and plotted 3D and 2D diagrams of the fluctuation function in
Figures 1 and 2, respectively.

7
Fig. 1. 3D diagrams of the fluctuation function for different numbers of biophotons N. It is seen that the maximum of fluctuation is around
Z=1. MT is initially in the ground state

As introduced above, F(T,z) is the fluctuation function which determines the transition rate of
biomolecules between the coherent states and the ground state. When in coherent states, the
biophotons are absorbed via biomolecules and when in the ground state, the biophotons are absorbed
via the vacuum. In this model, the information can be restored from the vacuum, and conscious states
can be repeated as before. Then, the information can be sent back to the memory site again via
emission of biophotons. This cycle can be repeated an arbitrary number of times. We have plotted the
F(T,z) function for different values of its variables. According to 3D diagrams in Figure 1, we let the
values near 1 for Z, since the main fluctuations are in this area.

8
Fig. 2. 2D diagrams of the fluctuation function for different values of N and Z in different scaled time intervals. MT are initially in the
ground state.

Now, we investigate the behavior of the fluctuation function with the assumption that the MT is first in the
excited state. Here ce (0)  1 and c g (0)  0 . With the substitution of these coefficients we have

 e (t )   cn cos( t n  1) n (15)
n0

 g (t )  i  cn sin(t n  1) n  1 (16)
n 0
Continuing our calculations for the fluctuation function we find that
2

Nn  z 
F (T , z )   e  N 2 2
1  sin ( n  1T )  cos (e
2
n  1T )
n0 n!  

We have plotted 3D and 2D diagrams of the fluctuation function in Figures 3 and 4, respectively. As shown

9
before, the maximum of fluctuations is found to be around Z=1. Here, we see again these fluctuations around
Z=1.

Fig. 3. 3D diagrams of the fluctuation function for different values of biophotons N. In this interaction, MT is initially in the coherent state

And as before, we have plotted F(T,z) for this case for different values of its variables.

10
Fig. 4. 2D diagrams of the fluctuation function for different values of N and Z. MT is initially in the coherent state.

It is also worth stressing that centrioles and cilia, which are complex microtubular structures, are involved in
photoreceptor functions in single cell organisms and primitive visual systems. Cilia are also found in all retinal
rod and cone cells. The dimensions of centrioles and cilia are comparable to the wavelengths of visible and
infrared light (Hameroff and Tuszynski, 2004). In a series of studies spanning a period of some 25 years G
Albrecht-Buehler (AB) demonstrated that living cells possess a spatial orientation mechanism located in the
centriole (Albrecht-Buehler 1994, 1995, 1997). This is based on an intricate arrangement of microtubule
filaments in two sets of nine triplets each of which is perpendicular to the other. This arrangement provides the
cell with a primitive ‘‘eye” that allows it to locate the position of other cells within a two to three degree
accuracy in the azimuthal plane and with respect to the axis perpendicular to it (Albrecht-Buehler 1997). He
further showed that electromagnetic signals are the triggers for the cells’ repositioning. It is still largely a

11
mystery how the reception of electromagnetic radiation is accomplished by the centriole. Another mystery
related to these observations is the original electromagnetic radiation emitted by a living cell (Tuszynski and
Dixon, 2001).

7. Discussion

There is no doubt that EEG waves are deeply involved with the basic functioning of the brain but the
origin and the exact function of EEG has remained a mystery. The EEG waves associated with two
distant neurons are strongly correlated and this supports the view that EEG waves are related to the
properties of the brain as a coherent quantum system. It is not possible for a scalp EEG to determine
the activity within a single dendrite or neuron. Rather, a surface EEG reading is the summation of the
synchronous activity of thousands of neurons that have similar spatial orientation, radial to the scalp2.
Synaptic transmission and axonal transfer of nerve impulses are too slow to organize
coordinated activity in large areas of the central nervous system. Numerous observations confirm this
view (Reinis et al. 2005). The duration of a synaptic transmission is at least 0.5 ms, thus the
transmission across thousands of synapses takes about hundreds or even thousands of milliseconds.
The transmission speed of action potentials varies between 0.5 m/s and 120 m/s along an axon. More
than 50% of the nerves fibers in the corpus callosum are without myelin, thus their speed is reduced to
0.5 m/s. How can these low velocities (i.e. classical signals) explain the fast processing in the nervous
system? We believe that quantum theory is able to explain some of the above mysteries. As an
example, recently it has been shown theoretically that the biological brain has the possibility to
achieve large quantum bit computing at room temperature, superior when compared with the
conventional processors (Musha, 2009).
Neuroscientists or brain specialists record the EEG diagrams of their patients when their eyes
are closed, because when their eyes are open the amplitudes of diagrams dramatically would be
reduced and they cannot diagnose changes in the amplitudes. In Figure 5 we represent the EEG
diagrams of a person with different situations as explained in the caption to the figure. When the eyes
are opened, the number of interacting photons with the biomolecules becomes high and the
amplitudes of EEG become low. It is clear when the intensity of incident light is high; it must produce
more action potentials than low intensity light. In classical physics, it is argued that this situation is
because the superposition of several waves tends to decrease the amplitude of the total wave. This
argument ignores the intensity of the incident light entering the eye. This classical argument also
cannot explain when and how the synchronization happens in order to decrease the amplitude of EEG.

Fig. 5. A schematic alpha-EEG diagram of a person when (a) the eyes are opened and then he/she closes the eyes (b) the eyes are closed and
then she/he opens his eyes and then closes them again.

2
Updated from : http://www.reference.com/browse/electro-encephalogram, 13 November 2009.

12
Our argument is based on the interaction of photons and biomolecules inside the neurons. When the
eyes start to be opened, the incident numbers of photons into the eyes increases and according to our
diagrams in previous sections the amplitudes are reduced. So far we have not been able to find an
exact relation between the EEG diagrams and the fluctuation function, but the synchronous and
coherent vibrations of billions electric dipoles of biomolecules cannot be ignored in the EEG
diagrams. MTs are particularly numerous in the brain where they form highly ordered bundles and are
the best candidate for long coherence and large synchrony (Hameroff and Penrose, 1996). The
argument for connection between Alpha-EEG diagrams and MTs activity is their similar behavior in
increasing and decreasing of amplitudes of fluctuation function for MTs and potential difference in
EEG in response to the intensity of photons. This similarity during opening and shutting of the eyes
indicates a significant relation between the EEG diagrams and the fluctuation function. Obtaining a
mathematical relation between fluctuation function in MTs and potential difference in neurons can be
an interesting subject of our future research.

The amplitude increment of EEG is probable after a small time after the opening of the eyes, because
the pupil reflects against the sudden light and forbids the photon increment and reduces photon
numbers. Thus, the number of interacting photons will be reduced.

Acknowledgments
Vahid Salari acknowledges S. B. University of Kerman for their support to research in this context. Majid
Rahnama thanks the Kerman Neuroscience Research Center (KNRC) for supporting this work, and also István
Bókkon gratefully acknowledges the support he has received from the BioLabor (www.biolabor.org), (Hungary)
during this research. His URL: http://bokkon-brain-imagery.5mp.eu. Jack Tuszynski acknowledges support
from NSERC (Canada) for his research. Michal Cifra acknowledges financial support from Czech Science
Foundation, P102/10/P454.

References:

Albrecht-Buehler G., 1994. Cellular infrared detector appears to be contained in the centrosome, Cell Motil. Cytoskeleton
27, 262–271.

Albrecht-Buehler G, 1995. Changes of cell behavior by near-infrared signals, Cell Motil. Cytoskeleton 32, 299–304.

Albrecht-Buehler G, 1997. Autofluorescence of live purple bacteria in the near infrared, Exp. Cell Res. 236, 43–50.

Alfinito E, Manka R, Vitiello G. 2000. Vacuum structure for expanding geometry. Class. Quantum Grav. 17, 93-111.

Arai M, Okada N. 2001. Vacuum structure of spontaneously broken N=2 supersymmetric gauge theory. Phys. Rev. D. 64,
025024.

Adamo AM, Llesuy SF, Pasquini JM, Boveris A. 1989. Brain chemiluminescence and oxidative stress in hyperthyroid rats.
Biochem. J. 263, 273-277.

Batyanov AP. 1984. Distant optical interaction of mitochondria. Bull. Exp. Biol. Med. 97, 740-742

Bischof M. 2005. Biophotons-The light in our cells. Journal of Optometric Phototherapy. 1-5.

Bókkon I. 2003. Creative Information. Journal of Biological Systems 1, 1-17.

Bókkon I. 2005. Dreams and Neuroholography: An Interdisciplinary Interpretation of Development of Homeotherm State in
Evolution. Sleep Hypnosis 7, 61-76.

Bókkon I. 2009. Visual perception and imagery: a new hypothesis. BioSystems 96, 178-184.

Bókkon I. 2008. Phosphene phenomenon: A new concept. BioSystems 92, 168-174.

13
Bohm D, Hiley, B.J. 1993. The undivided universe. New York: Routledge,.

Bajpai, R. P. 1998. Coherent nature of biophotons: experimental evidence and phenomenological model. In Chang, J.-J.,
Fisch, J. Popp, F.-A. (Eds.), Biophotons.

Bajpai, R.P. 1999. Coherent Nature of the Radiation Emitted in Delayed Luminescence of Leaves. J. Theoretical Biol. 198,
287-299.

Bókkon I, Salari V, Tuszynski J, Antal I, 2010. Estimation of the number of biophotons involved in the visual perception of
a single object image: Biophoton intensity can be considerably higher inside cells than outside. Journal of Photochemistry
and Photobiology B: Biology 100, 160-166.

Chwirot BW. 1992. Ultraweak luminescence studies of microsporogenesis in Larch. pp. 259-285. In: Recent advances in
biophoton research and its applications. (Ed) by Popp FA, Li KH, Gu Q. Singapore ; River Edge, N.J. : World Scientific,.

Cifra M, Pokorný J, Havelka D, Kučera O. 2010. Electric field generated by axial longitudinal vibration modes of
microtubule. BioSystems, 100, 122 – 131.

Cilento G. 1988. Photobiochemistry without light. Experientia 44, 572-576.

Chiarugi P. 2003. Reactive oxygen species as mediators of cell adhesion. Ital. J. Biochem. 52, 28-32.

Chiarugi P. 2008. Src redox regulation: there is more than meets the eye. Mol. Cells. 26, 329-337.

Clegg J.S. 1984. Properties and metabolism of the aqueous cytoplasm and its boundaries. Amer. J. Physiol. 246, R133-R151

Clegg, J.S. 1981. Coll. Phenomena. 3, 289.

Deriu MA, Soncini M, Orsi M, Patel M, Essex JW, Montevecchi FM, Redaelli A. 2010. Anisotropic Elastic Network
Modeling of Entire Microtubules. Biophys. J. 99, 2190 – 2199.

Dröge W. 2002. Free Radicals in the Physiological Control of Cell Function. Physiol. Rev. 82, 47-95.

Faber J, Portugal R, Rosa LP. 2006. Information processing in brain microtubules. BioSystems 83, 1-9.

Frieden B.R. 1998. Physics from Fisher Information. Cambridge: Cambridge University Press.

Fröhlich H. 1968. Long range coherence and energy storage in biological systems. Int. J. Quantum Chem. 2, 641-649.

Fröhlich H. 1970. Long range coherence and the action of enzymes. Nature 228, 1093.

Fröhlich H. 1975. The extraordinary dielectric properties of biological materials and the action of enzymes. PNAS 72, 4211-
4215.

Grass F, Kasper S. 2008. Humoral phototransduction: light transportation in the blood, and possible biological effects. Med.
Hypotheses 71, 314-317.

Gordeeva AV, Zvyagilskaya RA, Labas YA. 2003. Cross-talk between reactive oxygen species and calcium in living cells.
Biochemistry (Mosc). 68, 1077-1080.

Gerry C., Knight P. 2005. Introductory Quantum Optics. Cambridge University Press.

Gu Qiao. 1999. On Coherence Theory of Biophoton Emission, Journal of the GCPD e.V. 5, 17-20.

Hameroff, S.R., Penrose, R. 1996. Conscious events as orchestrated space-time selections. J. Conscious. Stud. 3, 36-53.

Hameroff, S., Tuszynski, J. 2004. Quantum states in proteins and protein assemblies: The essence of life? Proceedings of
SPIE Conference on Fluctuationa and Noise, Canary Islands.

Hagan S, Hameroff S.R. Tuszynski J.A. 2002. Quantum computation in brain microtubules: Decoherence and biological
feasibility. Phys Rev E 65, 061901.
Hoidal JR. 2001. Reactive Oxygen Species and Cell Signaling Am. J. Respir. Cell Mol. Biol. 25, 661-663.

14
Hidalgo C, Carrasco MA, Muñoz P, Núñez MT. 2007. A role for reactive oxygen/nitrogen species and iron on neuronal
synaptic plasticity. Antioxid Redox Signal. 9, 245-255.

Hideg E, Kobayashi M, Inaba H. 1991. Spontaneous ultraweak light emission from respiring spinach leaf mitochondria.
BBA-Bioenerg. 1098, 27-31

Imaizumi S, Kayama T, Suzuki J. 1984. Chemiluminescence in hypoxic brain--the first report. Correlation between energy
metabolism and free radical reaction. Stroke. 15, 1061-1065.

Isojima Y, Isoshima T, Nagai K, Kikuchi K, Nakagawa H. 1995. Ultraweak biochemiluminescence detected from rat
hippocampal slices. NeuroReport 6, 658-660.

Jelínek F, Pokorný J. 2001. Microtubules in biological cells as circular waveguides and resonators. Electro- magnetobiol.,
20, 75-80.

Jibu M, Hagan S, Hameroff SR, Pribram KH, Yasue K. 1994. Quantum optical coherence in cytoskeletal microtubules:
implications for brain function. Biosystems 32, 195-209.

Jibu, M., Brinbram, K.H. Yasue, K. 1996. From conscious Experience to Memory Storage and Retrieval: The role of
quantum brain dynamics and boson condensation of evanescent photons. Inter. J. Mod. Physics. B. 10, 1735-1754.

Jaynes E.T. Cummings F. W. 1963. Proc. IEEE 51, 89.

Jibu, M. Yasue, K. 1995. Quantum brain dynamics: an introduction. Amsterdam: John Benjamins.

Kataoka Y, Cui Y, Yamagata A, Niigaki M, Hirohata T, Oishi N, Watanabe Y. 2001. Activity-Dependent Neural Tissue
Oxidation Emits Intrinsic Ultraweak Photons. Biochem Biophys Res Commun. 285, 1007-1011.

Kato M, Shinzawa K, Yoshikawa S. 1981. Cytochrome oxidase is a possible photoreceptor in mitochondria. JPhotochem
Photobiol. 2, 263-269.

Karu T. 1999. Primary and secondary mechanisms of action of visible to near-IR radiation on cells. J. Photochem. Photobiol.
B 49, 1-17.

Kobayashi M, Takeda M, Ito K, Kato H, Inaba H. 1999. Two-dimensional photon counting imaging and spatiotemporal
characterization of ultraweak photon emission from a rat's brain in vivo. J. Neurosci. Methods. 93, 163-168.

Kobayashi M, Takeda M, Sato T, Yamazaki Y, Kaneko K, Ito K, Kato H, Inaba H. 1999. In vivo imaging of spontaneous
ultraweak photon emission from a rat’s brain correlated with cerebral energy metabolism and oxidative stress. Neurosci. Res.
34, 103-113.

Kruk I, Lichszteld K, Michalska T, Wronska J, Bounias M. 1989. The formation of singlet oxygen during oxidation of
catechol amines as detected by infrared chemiluminescence and spectrophotometric method. Z Naturforsch. [C]. 44, 895-
900.

Kishida KT, Klann E. 2007. Sources and targets of reactive oxygen species in synaptic plasticity and memory. Antioxid.
Redox Signal. 9, 233-244.

Knapp LT, Klann E. 2002. Role of reactive oxygen species in hippocampal long-term potentiation: contributory or
inhibitory? J Neurosci Res. 70, 1-7.

Kalmus, Peter I P. 1999. Empty matter and the full physical vacuum. Phys. Educ. 34, 205-208.

Lifshits M.A. 1972. Biofiz. 17, 694.

Mansfield JW. 2005. Biophoton distress flares signal the onset of the hypersensitive reaction. Trends Plant Sci. 10, 307-309.

Mavromatos, N.E. , Mershin, A., Nanopoulos, D.V. 2002. QED-cavity model of microtubules implies dissipationless energy
transfer and biological quantum teleportation. Inter. J. Mod. Physics. B 16, 3623-3642.

Mavromatos N.E. Nanopoulos D.V. 1998. On Quantum Mechanical Aspects of Microtubules. Inter. J. Mod. Physics. B 12,
517 -542.

15
Murata T, Kadota A, Wada M. 1997. Effects of Blue Light on Cell Elongation and Microtubule Orientation in Dark-Grown
Gametophytes of Ceratopteris richardii. Plant. Cell. Physiol. 38, 201-209.

Mazhul' VM, Shcherbin DG. 1999. Phosphorescent analysis of lipid peroxidation products in liposomes. Biofizika 44, 676-
681.

Musha T. 2009. Possibility of High Performance Quantum Computation by Superluminal Evanescent Photons in Living
Systems. BioSystems 96, 242-245.

Matsuno, K, Paton, R.C. 2000. Is there biology of quantum information? BioSystems 55, 39-46.

Nakano M: 1989. Low-level chemiluminescence during lipid peroxidations and enzymatic reactions. J. Biolumin.
Chemilum. 4, 231-240.

Nogales E, Wolf SG,. Downing K. 1998. Structure of the alpha beta tubulin dimer by electron crystallography. Nature, 391,
199-203.

Ouyang, M., Awschalom, D. D. 2003. coherent spin transfer between moleculary bridged quantum dots. Science, 301, 1074-
1078.

Pokorný J, Jelínek F, Trkal V, Lamprecht I, Hölzel R. 1997, Vibrations in Microtubules. J. Biol. Phys. 23, 171-179.

Popp F.A. 1979. Photon storage in biological systems. in Popp, F.A., Becker, G., König, H.L. Peschka, W. (Eds.),
Electromagnetic bio-information.

Popp FA, Chang JJ, Herzog A, Yan Z, Yan Y. 2002. Evidence of non-classical (squeezed) light in biological systems. Phys.
Lett. A 293, 98-102.

Popp FA, Yan Y. 2002. Delayed Luminescence of Biological Systems in Terms of Coherent States. Phys. Lett. A 293, 93-
97.

Popp FA, Nagl W, Li KH, Scholz W, Weingartner O, Wolf R. 1984. Biophoton emission. New evidence for coherence and
DNA as source. Cell. Biophys. 6, 33-52.

Plaga. R. 2006. New physics beyond the standard model of particle physics and parallel universes. Physics Lett. B. 634, 116-
118.

Quickenden TI., Que Hee, SS: 1974. Weak luminescence from the yeast Sachharomyces-Cervisiae. Biochem Biophys Res
Commun. 60, 764-770.

Reimers J.R. McKemmish L.K. McKenzie R.H. Mark A.E. Hush N.S. 2009. PNAS 0806273106, doi/10.1073.

Reinis S, Holub RF, Smrz P. 2005. A quantum hypothesis of brain function and consciousness,translated version of
Ceskoslovenska fyziologie 54 .26-31.

Rosa, L., P., Faber, J., 2004. Quantum Models of the Mind: Are they compatible with environment decoherence?, Physical
Review E, 70, 031902.

Salari V, Tuszynski J, Rahnama M, Bernroider G. 2010. On the Possibility of Quantum Coherence in Biological Systems,
DICE 2010, to appear in Journal of Physics, Conf. Ser. (2011).

Salari V, Rahnama M, Tuszynski J. 2011. Toward Quantum Visual Information Transfer in the Human Brain. To appear in
Foundations of Physics..

Skulachev V.P. 2001. Mitochondrial filaments and clusters as intracellular power-transmitting cables. Trends Biochem. Scie.
26, 23-29.

Scott RQ, Roschger P, Devaraj B, Inaba H. 1991. Monitoring a mammalian nuclear membrane phase transition by intrinsic
ultraweak light emission. FEBS Lett. 285, 97-98.

16
Slawinski J. 1988. Luminescence research and its relation to ultraweak. cell radiation. Experientia 44, 559-571.

Stonier T. 1990. Information and the Internal Structure of the Universe. An Exploration into Information Physics, Berlin:
Springer.

Staxén I, Bergounioux C, Bornman J. 1993. Effect of ultraviolet radiation on cell division and microtubule organization in
Petunia hybrida protoplasts. Protoplasma, 173, 70-76

Sun Y, Wang C, Dai J. 2010. Biophotons as neural communication signals demonstrated by in situ biophoton autography.
Photochem. Photobiol. Sci. 9, 315-322

Thar R, Kühl M. 2004. Propagation of electromagnetic radiation in mitochondria? J. Theor. Biol. 230, 261-270.

Tilbury RN, Cluickenden TI. 1988. Spectral and time dependence studies of the ultraweak bioluminescence emitted by the
bacterium Escherichia coli. Photobiochem Photobiophys. 47, 145-150.

Takeda M, Tanno Y, Kobayashi M et al. 1998. A novel method of assessing carcinoma cell proliferation by biophoton
emission. Cancer Lett. 127, 155-60.

Touyz RM. 2005. Reactive oxygen species as mediators of calcium signaling by angiotensin II: implications in vascular
physiology and pathophysiology. Antioxid. Redox Signal. 7, 1302-1314.

Tejada-Simon MV, Serrano F, Villasana LE, Kanterewicz BI, Wu GY, Quinn MT, Klann E. 2005. Synaptic localization of a
functional NADPH oxidase in the mouse hippocampus. Mol. Cell. Neurosci. 29, 97-106.

Thiels E, Urban NN, Gonzalez-Burgos GR, Kanterewicz BI, Barrionuevo G, Chu CT, Oury TD, Klann E. 2000. Impairment
of long-term potentiation and associative memory in mice that overexpress extracellular superoxide dismutase. J Neurosci.
20, 7631-7639.

Thiels E, Klann E. 2002. Hippocampal memory and plasticity in superoxide dismutase mutant mice. Physiol. Behav.
77, 601-605.

Tong J.J. 2007. Mitochondrial delivery is essential for synaptic potentiation. The Biological Bulletin 212, 169-175.

Turing, A.M. 1950. Computing machinery and intelligence. Mind 59, 433-460.

Timothy H.B. 1985. The classical vacuum (zero-point energy). Scientific American Aug. 70-78.

Tegmark M. 2000. Importance of quantum decoherence in brain processes. Phys Rev E 61, 4194–4206.

Tuszynski JA, Dixon JM, 2001. Quantitative analysis of the frequency spectrum of the radiation emitted by cytochrome
oxidase enzymes, Phys. Rev. E 64 051915.

Van Wijk R, Schamhart DH. 1988. Regulatory aspects of low intensity photon emission. Experientia. 44, 586-593.

Van Wijk R., Bosman S., Ackerman J., Van Wijk E. P. A. 2008. NeuroQuantology l6, 452-463.

Valko M, Leibfritz D, Moncol J, Cronin MT, Mazur M, Telser J: 2007. Free radicals and antioxidants in normal
physiological functions and human disease. Int. J. Biochem. Cell. Biol. 39, 44–84

Vladimirov YA, Proskurnina EV. 2009. Free radicals and cell chemiluminescence. Biochem. (Moscow), 74, 1545-1566,
original txt in Uspekhi Biologicheskoi Khimii. 2009. 49, 341-388.

Volterra A, Trotti D, Tromba C, Floridi S, Racagni G. 1994. Glutamate uptake inhibition by oxygen free radicals in rat
cortical astrocytes. J. Neurosci. 14, 2924-2932

Watts BP, Barnard M, Turrens JF, 1995. Peroxynitrite-Dependent Chemiluminescence of Amino Acids, Proteins, and Intact
Cells. Arch. Biochem. Biophys. 317, 324-330.

Wu T.M. Austin S. 1977. Bose condensation in biosystems. Phys Lett A 64, 151–152.

Wu T.M. Austin S. 1978. Cooperative behavior in biological systems. Phys Lett A 65, 74–76.

17
Wu TM Austin SJ 1981. Froehlich’s model of Bose condensation in biological systems. J Bio Phys 9, 97–107.

Yoon YZ, Kim J, Lee BC, Kim YU, Lee SK, Soh KS. 2005. Changes in ultraweak photon emission and heart rate variability
of epinephrine-injected rats. Gen Physiol Biophys. 24, 147–159.

Wang Ch, Bókkon I, Dai J, Antal I. 2010. Spontaneous and visible light-induced ultraweak photon emission from rat eyes.
Brain Res. DOI: 10.1016/j.brainres.2010.10.077.

18

Anda mungkin juga menyukai