ChapterPDF Available

The role of nitric oxide in low level light therapy

Authors:

Abstract and Figures

The use of low levels of visible or near infrared light for reducing pain, inflammation and edema, promoting healing of wounds, deeper tissues and nerves, and preventing tissue damage by reducing cellular apoptosis has been known for almost forty years since the invention of lasers. Despite many reports of positive findings from experiments conducted in vitro, in animal models and in randomized controlled clinical trials, LLLT remains controversial. Firstly the biochemical mechanisms underlying the positive effects are incompletely understood, and secondly the complexity of choosing amongst a large number of illumination parameters has led to the publication of a number of negative studies as well as many positive ones. This review will focus on the role of nitric oxide in the cellular and tissue effects of LLLT. Red and near-IR light is primarily absorbed by cytochrome c oxidase (unit four in the mitochondrial respiratory chain). Nitric oxide produced in the mitochondria can inhibit respiration by binding to cytochrome c oxidase and competitively displacing oxygen, especially in stressed or hypoxic cells. If light absorption displaced the nitric oxide and thus allowed the cytochrome c oxidase to recover and cellular respiration to resume, this would explain many of the observations made in LLLT. Why the effect is only seen in hypoxic, stressed or damaged cells or tissues? How the effects can keep working for some time (hours or days) postillumination? Why increased NO concentrations are sometimes measured in cell culture or in animals? How blood flow can be increased? Why angiogenesis is sometimes increased after LLLT in vivo?
Content may be subject to copyright.
The Role of Nitric Oxide in Low Level Light Therapy.
Michael R Hamblin a,b,c,*
a Wellman Center for Photomedicine, Massachusetts General Hospital, b Department of Dermatology, Harvard
Medical School, c Harvard-MIT Division of Health Sciences and Technology,
ABSTRACT
The use of low levels of visible or near infrared light for reducing pain, inflammation and edema, promoting healing
of wounds, deeper tissues and nerves, and preventing tissue damage by reducing cellular apoptosis has been known
for almost forty years since the invention of lasers. Despite many reports of positive findings from experiments
conducted in vitro, in animal models and in randomized controlled clinical trials, LLLT remains controversial.
Firstly the biochemical mechanisms underlying the positive effects are incompletely understood, and secondly the
complexity of choosing amongst a large number of illumination parameters has led to the publication of a number of
negative studies as well as many positive ones. This review will focus on the role of nitric oxide in the cellular and
tissue effects of LLLT. Red and near-IR light is primarily absorbed by cytochrome c oxidase (unit four in the
mitochondrial respiratory chain). Nitric oxide produced in the mitochondria can inhibit respiration by binding to
cytochrome c oxidase and competitively displacing oxygen, especially in stressed or hypoxic cells. If light
absorption displaced the nitric oxide and thus allowed the cytochrome c oxidase to recover and cellular respiration to
resume, this would explain many of the observations made in LLLT. Why the effect is only seen in hypoxic,
stressed or damaged cells or tissues? How the effects can keep working for some time (hours or days) post-
illumination? Why increased NO concentrations are sometimes measured in cell culture or in animals? How blood
flow can be increased? Why angiogenesis is sometimes increased after LLLT in vivo?
Keywords: biostimulation, low level laser therapy, mitochondria, cytochrome c oxidase, nitric oxide
1. LOW LEVEL LIGHT THERAPY
Although low level light therapy (LLLT) has been known and increasingly widely practiced for over forty
years, it is still regarded with some skepticism by laymen and medical professionals alike, and has not reached
acceptance by mainstream medicine. The single most important reason for this lack of acceptance is likely to be the
inability of most practitioners of LLLT to satisfactorily explain how it works on a molecular, cellular and tissue
level. There is a need for more fundamental research on identifying photoacceptor molecules, elucidating cell and
signaling pathways that are engaged after cells absorb visible photons. Furthermore it is necessary to investigating
relationships between the optical parameters of the light such as wavelength, total delivered energy, rate at which
energy is delivered, coherence, polarization state and pulse structure
Figure 1. Schematic representation of the main areas of application of LLLT
Cellular
photoreceptor
Cellular
photoreceptor
Wound healing
Tissue repair
Prevention of tissue death
Relief of inflammation
Pain, edema
Acute injuries
Chronic diseases
Neurogenic pain
Neurological problems
Acupuncture
hν, 600-950-nm,
Invited Paper
Mechanisms for Low-Light Therapy III, edited by Michael R. Hamblin, Ronald W. Waynant,
Juanita Anders, Proc. of SPIE Vol. 6846, 684602, (2008) · 1605-7422/08/$18 · doi: 10.1117/12.764918
Proc. of SPIE Vol. 6846 684602-1
2008 SPIE Digital Library -- Subscriber Archive Copy
inter membrais space
P,bosome cristae
Granus\
(NH'C
ATP synnhase pmniclne
Inns, membrane
O,Ier membfane
2. MITOCHONDRIA CONTAIN CELLULAR PHOTOACCEPTORS
2.1 Mitochondria
Mitochondria are distinct organelles with two membranes and are usually rod-shaped. Mitochondria are sometimes
described as "cellular power plants," because they convert food molecules into energy in the form of ATP via the
process of oxidative phosphorylation. A typical eukaryotic cell contains about 2,000 mitochondria, which occupy
roughly one fifth of its total volume. Mitochondria contain DNA that is independent of the DNA located in the cell
nucleus. Mitochondrial DNA is circular and lies in the matrix in punctate structures called "nucleoids" each
containing 4-5 copies of the mitochondrial DNA (mtDNA). Mitochondria have their own ribosomes, and can make
many of their own proteins. The outer membrane limits the organelle, while the inner membrane is thrown into folds
or shelves that project inward called "cristae mitochondriales". A mitochondrion contains inner and outer
membranes composed of phospholipid bilayers and proteins, and consequently there are 5 distinct compartments
within mitochondria. There is the outer membrane, the intermembrane space (the space between the outer and inner
membranes), the inner membrane, the cristae space (formed by infoldings of the inner membrane), and the matrix
(space within the inner membrane). Mitochondria range from 1 to 10 µm in size.
Figure 2. Schematic representation of the structure of a mitochondrion in a mammalian cell
A dominant role for the mitochondria is the production of ATP as reflected by the large number of proteins in the
inner membrane needed for this task. This is done by oxidizing the major products of glycolysis, pyruvate and
NADH that are produced in the cytosol. This process of cellular respiration, also known as aerobic respiration, is
dependent on the presence of oxygen. When oxygen is limited the glycolytic products will be metabolized by
anaerobic respiration, a process that is independent of the mitochondria. The production of ATP from glucose has an
approximately 15-fold higher yield during aerobic respiration compared to anaerobic respiration. In addition to their
role in producing cellular energy in the form of ATP, mitochondria play an important role in many other metabolic
tasks, such as, apoptosis (programmed cell death), glutamate-mediated excitotoxic neuronal injury, cellular
proliferation, regulation of the cellular redox state, heme synthesis and steroid synthesis.
Each pyruvate molecule produced by glycolysis is actively transported across the inner mitochondrial membrane,
and into the matrix where it is oxidized and combined with coenzyme A to form CO2, acetyl-CoA and NADH. The
acetyl-CoA is the primary substrate to enter the citric acid cycle, also known as the tricarboxylic acid (TCA) cycle
or Krebs cycle. The enzymes of the citric acid cycle are located in the mitochondrial matrix with the exception of
succinate dehydrogenase, which is bound to the inner mitochondrial membrane. The citric acid cycle oxidizes the
acetyl-CoA to carbon dioxide and in the process produces reduced cofactors (three molecules of NADH and one
molecule of FADH2), that are a source of electrons for the electron transport chain, and a molecule of GTP (that is
readily converted to ATP).
Proc. of SPIE Vol. 6846 684602-2
2.2 Mitochondrial Respiratory Chain
Four membrane-bound complexes have been identified in mitochondria. Each is an extremely complex
transmembrane structure that is embedded in the inner membrane. Three of them are proton pumps. The structures
are electrically connected by lipid-soluble electron carriers and water-soluble electron carriers.
Figure 3. Structure of the electron transport chain in the mitochondrial inner membrane
2.2.1 Complex I
Complex I (NADH dehydrogenase, also called NADH:ubiquinone oxidoreductase; EC 1.6.5.3) removes two
electrons from NADH and transfers them to a lipid-soluble carrier, ubiquinone (Q). The reduced product, ubiquinol
(QH2) is free to diffuse within the membrane. At the same time, Complex I moves four protons (H+) across the
membrane, producing a proton gradient. Complex I is one of the main sites at which premature electron leakage to
oxygen occurs, thus being one of main sites of production of a harmful free radical called superoxide.
NADH is oxidized to NAD+, reducing Flavin mononucleotide to FMNH2 in one two-electron step. The next
electron carrier is a Fe-S cluster, which can only accept one electron at a time to reduce the ferric ion into a ferrous
ion. In a convenient manner, FMNH2 can be oxidized in only two one-electron steps, through a semiquinone
intermediate. The electron thus travels from the FMNH2 to the Fe-S cluster, then from the Fe-S cluster to the
oxidized Q to give the free-radical (semiquinone) form of Q. This happens again to reduce the semiquinone form to
the ubiquinol form, QH2. During this process, four protons are translocated across the inner mitochondrial
membrane, from the matrix to the intermembrane space. This creates a proton gradient that will be later used to
generate ATP through oxidative phosphorylation.
2.2.2 Complex II
Complex II (succinate dehydrogenase; EC 1.3.5.1) is not a proton pump. It serves to funnel additional electrons into
the quinone pool (Q) by removing electrons from succinate and transferring them (via FAD) to Q. Complex II
consists of four protein subunits: SDHA,SDHB,SDHC, and SDHD. Other electron donors (e.g., fatty acids and
glycerol 3-phosphate) also funnel electrons into Q (via FAD), again without producing a proton gradient.
2.2.3 Complex III
Complex III (cytochrome bc1 complex; EC 1.10.2.2) removes in a stepwise fashion two electrons from QH2 and
transfers them to two molecules of cytochrome c, a water-soluble electron carrier located within the intermembrane
space. At the same time, it moves two protons across the membrane, producing a proton gradient (in total 4 protons:
2 protons are translocated and 2 protons are released from ubiquinol). When electron transfer is hindered (by a high
+ + + + + + + + + + + + + + +
Intermembrane
Intermembrane
space
space
Mitochondrial
Mitochondrial
matrix
matrix
Cyto C2+
III
H+
ADP+PiATP
4H+2H+
4H+
I
I
NADH2
NAD+
succinate fumarate
II Q
QH2
1/2O2+2e-
H2O
- - - - - - -
e-
e-
IV
Cyto C3+
e-
Proc. of SPIE Vol. 6846 684602-3
membrane potential, point mutations or respiratory inhibitors such as antimycin A), Complex III may leak electrons
to oxygen resulting in the formation of superoxide, a highly-toxic species, which is thought to contribute to the
pathology of a number of diseases, including aging.
2.2.4 Complex IV
Complex IV (cytochrome c oxidase; EC 1.9.3.1) removes four electrons from four molecules of cytochrome c and
transfers them to molecular oxygen (O2), producing two molecules of water (H2O). At the same time, it moves four
protons across the membrane, producing a proton gradient.
2.3 Mitochondria absorb visible light.
Several pieces of evidence suggest that mitochondria are responsible for the cellular response to red visible and NIR
light. The most popular system to study is the effects of HeNe laser illumination of mitochondria isolated from rat
liver. Increased proton electrochemical potential and ATP synthesis was found [1]. Increased RNA and protein
synthesis was demonstrated after 5 J/cm2 of HeNe laser light [2]. Pastore et al [3] found increased activity of
cytochrome c oxidase and an increase in polarographically measured oxygen uptake after 2 J/cm2 of HeNe laser. A
major stimulation in the proton pumping activity, about 55% increase of <--H+/e- ratio was found in illuminated
mitochondria. Yu et al [4] used 660 nm laser at a power density of 10 mW/cm2 and showed increased oxygen
consumption (0.6 J/cm2 and 1.2 J/cm2), increased phosphate potential, and energy charge (1.8 J/cm2 and 2.4 J/cm2)
and enhanced activities of NADH: ubiquinone oxidoreductase, ubiquinol: ferricytochrome c oxidoreductase and
ferrocytochrome C: oxygen oxidoreductase (between 0.6 J/cm2, and 4.8 J/cm2).
Irradiation of mitochondria with light at wavelengths of 650, and 725 nm [5] enhanced ATP synthesis. Light at
wavelengths of 477 and 554 nm did not influence the rate of this process. Oxygen consumption was increased by
illuminating with light at 365 and 436 nm, but not at 313, 546, and 577 nm [6]. Irradiation with light at 633 nm
increased the mitochondrial membrane potential and proton gradient, caused changes in mitochondrial optical
properties, modified some NADH-linked dehydrogenase reactions (NADH is a reduced form of nicotinamide
adenine dinucleotide), and increased the rate of ADP/ATP exchange (ADP is adenosine diphosphate) [7], as well as
RNA and protein synthesis in the mitochondria. In the case of state 4 respiration, 351 and 458 nm laser irradiations
accelerated the oxygen consumption of rat liver mitochondria; such an acceleration was not observed with 514.5 nm
irradiation. In the case of state 4 respiration (slower rate after all the ADP has been phosphorylated to form ATP),
351 and 458 nm laser irradiations accelerated the oxygen consumption of rat liver mitochondria; such an
acceleration was not observed with 514.5 nm irradiation. On the contrary, in the case of state 3 respiration (active
rate in presence of sufficient substrate, O2 and ADP), 514.5 nm argon laser irradiation activated the oxygen
consumption of mitochondria. Activation did not occur with 458 nm irradiation and 351 nm irradiation reduced the
oxygen consumption in state 3 [8]. 660 nm irradiation increased state 3 oxygen consumption, as well as increasing
the respiratory control ratio [4]. It is also believed that mitochondria are the primary targets when the whole cells are
irradiated with light at 630, 632.8 [9-11], or 820 nm. Irradiation with light at 812 [12] or 632.8 nm altered the
rhodamine 123 uptake by fibroblasts. These results were interpreted by the authors as inducing the perturbation of
mitochondrial energy production and membrane potential.
2.3 Cytochrome c oxidase is a photoacceptor.
In 1995, Karu defined the action spectra for mammalian cells of several processes stimulated by LLLT such as
DNA and RNA synthesis, and cellular adhesion [13]. The action spectra for all of these secondary markers were
very similar suggesting a common photoacceptor that can transduce light energy to accelerate all these processes.
Karu then compared these action spectra with visible and NIR absorption spectra of the copper centers of
cytochrome c oxidase in both reduced and oxidized states. Cytochrome c oxidase contains four redox active metal
centers and has a strong absorbance in the near infrared spectral range. The spectral absorbance of cytochrome c
oxidase and the action spectra were very similar. Based on this, Karu suggested that the primary photoacceptors are
mixed valence copper centers within cytochrome c oxidase [14]. Cytochrome c oxidase is the terminal enzyme of
the mitochondrial electron transport chain of all eukaryotes and is required for the proper function of almost all cells
especially those of highly metabolically active organs, such as the brain and heart. Recently, work from the Whelan
Proc. of SPIE Vol. 6846 684602-4
group from Medical College of Wisconsin has also suggested that cytochrome c is the critical chromophore
responsible for stimulatory effects of irradiation with infrared light [15-17]. Wong-Riley et al. [18] demonstrated
that infrared irradiation reversed the reduction in cytochrome c oxidase activity produced by the blockade of voltage
dependent sodium channels with tetrodotoxin and up regulated cytochrome c activity in primary neuronal cells . In
vivo Eells et al demonstrated that rat retinal neurons are protected from damage induced by methanol intoxication
[19]. The actual toxic metabolite formed from methanol is formic acid which inhibits cytochrome c.
Figure 4. Structure and mode of action of cytochrome c oxidase
Absorption spectra obtained for cytochrome c oxidase in different oxidation states were recorded and found to be
very similar to the action spectra for biological responses to light. Therefore it was proposed that cytochrome c
oxidase is the primary photoacceptor for the red-NIR range in mammalian cells [13]. Cytochrome c oxidase
(Structure is shown in Figure 4) contains two iron centers, haem a and haem a3 (also referred to as cytochromes a
and a3), and two copper centers, CuA and CuB [20] . Fully oxidized cytochrome c oxidase has both iron atoms in the
Fe(III) oxidation state and both copper atoms in the Cu(II) oxidation state, while fully reduced cytochrome c oxidase
has the iron in Fe(II) and copper in Cu(I) oxidation states. There are many intermediate mixed-valence forms of the
enzyme and other coordinate ligands such as CO, CN, and formate can be involved. All the many individual
oxidation states of the enzyme have different absorption spectra [21], thus probably accounting for slight differences
in action spectra of LLLT that have been reported. A recent paper from Karu’s group [14] gave the following
wavelength ranges for four peaks in the LLLT action spectrum: 1) 613.5 - 623.5 nm, 2) 667.5 - 683.7 nm, 3) 750.7 -
772.3 nm, 4) 812.5 - 846.0 nm.
A study from Pastore et al [22] examined the effect of He-Ne laser illumination on the purified cytochrome c
oxidase enzyme and found increased oxidation of cytochrome c and increased electron transfer. Artyukhov and
colleagues found [23] increased enzyme activity of a different enzyme catalase after He-Ne illumination. Absorption
of photons by molecules leads to electronically excited states and consequently can lead to acceleration of electron
transfer reactions [4]. More electron transport necessarily leads to increased production of ATP [24].
3. NITRIC OXIDE
3.1 Formation and action of NO.
O2+4 Cyt c2+out+8H+in 2H2O+4 Cyt c3+out+4H+out
Cyt C2+
H2O
CuBHaem a3
Glu
Asp
H+
H+
H+
H+
O
H+
H+
O
H+
H+
O
H+
H+
O
e-
e-
e-
e-
e-
e-
e-
e-
H+
O2
e-
e-
Cyt C3+
H+O2
Haem a
CuA
Intermembrane
Intermembrane
space
space
Mitochondrial
Mitochondrial
matrix
matrix
H+
Proc. of SPIE Vol. 6846 684602-5
Nitric oxide (NO), a free radical gas that is a powerful regulator of circulation (it is an endogenous vasodilator) and
a neurotransmitter (it helps in the processing of nerve signals as they cross synapses). L-arginine, one of 20 amino
acids that make up proteins, is the only amino acid that generates significant amounts of NO. The Nobel Prize was
awarded to three Americans in 1998 for their work on discovering NO and clarifying its role in health [25]. Their
most important contributions lay in describing the effect of NO on the circulation. The blood flow and nerve
responses are rapid. Small increases in NO lead to both vasodilation and to better sensory perception. NO
metabolism is necessary for normal circulation (venous, arterial, and lymph flows) and for the ability to sense pain,
temperature, and pressure.
The amino acid L-arginine, that is the main source of NO is released from proteins and small peptides in the small
intestine and is then absorbed, along with other amino acids into the circulation from which it is delivered to every
cell in the body. Some L-arginine is metabolized for NO synthesis and some is used for protein synthesis. In
endothelial cells, the small cells that make up capillaries and line every blood vessel and lymph duct in the body, L-
arginine can be converted to NO. This occurs only if the enzyme that makes NO and its co-factors are available in
adequate amounts. In diabetic patients and those with atherosclerotic disease plaques often occludes a portion of a
vessel so that the endothelial cells are not able to properly absorb NO. If the endothelial cell cannot take up L-
arginine, then NO synthesis will be impaired. Moreover, if atherosclerotic disease is present, oxygen delivery to all
cells is impaired and molecular oxygen is one of the cofactors needed by the enzyme to generate NO from L-
arginine. The NO diffuses into the smooth muscle cells that surround the endothelial lining of blood vessels cells
causing a biologic chain of events that lead to smooth muscle cell relaxation. This results in more blood flow to the
tissues. Tissues that are hypoxic (deprived of good, normal circulation) can not produce as much NO as do normal,
well oxygenated tissues. Thus an initial period of hypoxia leads to declines in NO production and less and less blood
flow over time.
Nitric oxide synthase (NOS) is the enzyme that generates NO from L-arginine. There are three different type of
NOS: neuronal nitric oxide synthase (NOS1), inducible nitric oxide synthase (NOS2) and endothelial nitric oxide
synthase (NOS3) [26]. Each of them have different tissue distributions and located on different human
chromosomes. They may related to many human diseases, such as Alzheimers dieseases, Parkinsons. diabetes,
asthma, heart disease, infection diseases. Often all three isoforms will be found in the same cell but occasionally one
cell will contain only one of the isoforms.
NOS1 is the neuronal (or brain) isoform. It helps in synaptic transmission, the processing of nervous information
from nerve to nerve, across gaps between the nerves called synapses, and from peripheral nerves to the brain.
NOS2 is called inducible or iNOS. This enzyme generates extraordinarily high concentrations of NO, in part to kill
bacteria. NOS2 (iNOS) takes several hours to be mobilized and the response is due to an injury or infectious
process. NOS2 produced by macrophages is responsible, in part, for their effects to repair injury and to ward off
infections. In other words, when the body mounts an inflammatory response to injury, macrophages are attracted to
the site of injury where they produce large amounts of NO. Extraordinarily high concentrations of NO (100 to 1000
times normal) are produced very locally by this isoform. In fact, reports suggest that wound (ulcer) fluid may
contain levels of NO that are very high and can only be attributed to iNOS. Unlike NOS1, which is part of normal
neurotransmission, there must be something very abnormal (a wound, tissue damage, hypoxia, bacterial infection,
etc.) to induce this enzyme.
The third isoform is eNOS (or NOS3) which stands for “endothelial” NOS. This isoform is active at all times (it
does not need to be induced as does iNOS) and is found in endothelial cells which are the cells that line the inner
surface of all blood vessels and lymph ducts. eNOS is activated by the pulsatile flow of blood through vessels. This
leads to a “shear stress” on the membrane of the endothelial cells as the column of blood in the vessel moves
forward and then stops. This NO, produced by eNOS, maintains the diameter of blood vessel so that perfusion of
tissues (skin, muscle, nerves, and bone) is maintained at optimal levels. In addition, eNOS mediated NO causes
angiogenesis, which is the growth of new blood vessels. This is especially important in healing an ulcer or wound on
the skin.
Proc. of SPIE Vol. 6846 684602-6
One interesting interplay of iNOS and eNOS is in tissue repair. Initially,NO is generated from iNOS in order to
ward off infection and to destroy and remove the irreversibly damaged, necrotic tissue. This is often referred to as
the inflammatory stage of wound repair. This phase lasts only a short time (a few days with an acute wound) and
then eNOS is (or should be) mobilized to cause vasodilation and angiogenesis to induce the healing response. NO
will relax smooth muscle cells and thus dilate veins, arteries, and lymphatics. This increases blood supply both to the
repairing tissues and from the damaged region. The latter removes metabolic waste products, reduces edema, and
prevents swelling that would otherwise compress capillaries. In the absence of adequate blood supply tissue will
remain hypoxic and heal only slowly, if at all. Moreover, since iNOS is produced in large part by white blood cells
(WBC), vasodilation permits delivery of additional WBC to the area that needs to be defended from infection. There
are wounds that do become infected and often only marginal reduction of the infection is seen even with high dose
and high potency antibiotics. If the vascular bed (arteries, veins, and lymphatics) were dilated, more of the antibiotic
would get to site of infection. Thus it is essential that eNOS be activated to produce NO. Clearly both eNOS and
iNOS play a role in wound healing; neither alone is sufficient to achieve full recovery. In diabetic patients, however,
eNOS activity is often well below normal so these patients cannot produce NO at normal levels.
3.2 Nitric oxide in mitochondria
Over the past decade it has been discovered that cells often use NO to block respiration. Nitric oxide is emitted by
nerve endings and can act on an enzyme called guanylate cyclase to relax blood vessels. For a long time, scientists
thought that guanylate cyclase was the only target of NO, but in the mid-1990s, they found that the molecule could
also bind to cytochrome oxidase and hinder respiration [27]. The finding that the body could poison one of its own
enzymes was initially shrugged off as an imperfection, but a few years later, several groups reported that
mitochondria contained a particular isoform of nitric oxide synthase [28]. This mitochondrial NOS was identified as
the neuronal isoform by Kanai et al [29]. Moncada proposed that evolution really crafted cytochrome oxidase to
bind not only oxygen but also NO [27]. One effect of slowing respiration in some locations would be to divert
oxygen elsewhere in cells and tissues. This prevents oxygen levels sinking dangerously low. NO blocks respiration
in the cells lining blood vessels and that this helps to transfer oxygen into smooth muscle cells in these vessels.
Respiration does not just generate energy, but it also generates feedback that allows a cell to monitor and respond to
its environment. Blocking respiration generates chemical signals, in the form of reactive oxygen species (ROS) such
as superoxide. ROS are normally associated with cell damage, but now it is thought they can interact with the
proteins that control gene activity and adapt cells to changing circumstances. In the past few years, researchers have
compiled a list of these proteins, or transcription factors, the activity of which depends , at least in part, on
interactions with ROS [30]. These include many proteins known to be linked to cellular life and death, such as p53.
The whole system is thought to controlled by cytochrome c oxidase, which catalyses the final step of respiration, in
which electrons and protons are transferred onto oxygen to form water. The cell can suppress the number of free
radicals coming from these respiratory chains by allowing protons to leak back though the membrane without
driving the synthesis of ATP, a process known as uncoupling [31]. But if uncoupling does not bring free-radical leak
under control, the signal may be amplified. Cells that depend on mitochondria for energy, such as neurons, may be pushed
to apoptosis by NO binding, making degenerative disease more likely. The activity of cytochrome c oxidase is inhibited
by nitric oxide (NO). This inhibition of mitochondrial respiration by NO can be explained by a direct competition
between NO and O2 for the reduced binuclear center CuB/a3 of cytochrome c oxidase and is reversible [32].
3.3 Interaction of Cytochrome C Oxidase With Nitric Oxide
The interaction of NO with cytochrome C oxidase in different types of cells is associated with the resistance to
apoptosis induced by various kinds of stressors, including growth factor deprivation [33], treatment with
staurosporine [34], O2 limitation [34], or intracellular calcium overload [35]. Depending on the system under study,
protection was shown to be associated with an increase in mitochondrial membrane potential (∆Ψm) [33], with an
increase in glycolytic output linked to upregulation of AMP-activated protein kinase (AMPK) [36], or with changes
in calcium efflux leading to the induction of the cytoprotective chaperone protein Grp78.[35]. Further studies also
showed that competition between NO and O2 at the level of cytochrome C oxidase is responsible for the inhibition
of hypoxia-inducible factor (HIF) 1-α stabilization observed in the presence of NO under otherwise limiting O2
concentrations [37], suggesting that mitochondria under the influence of NO may also be involved in the attenuation
of adaptive responses to low O2. In addition, there is evidence that NO promotes mitochondrial biogenesis by a
mechanism that is independent of cytochrome C oxidase but involves activation of the soluble guanylate cyclase.
Proc. of SPIE Vol. 6846 684602-7
Excessive production of NO and mitochondrial dysfunction have for many years been independently associated with
pathophysiological mechanisms. However, the fact that NO inhibits mitochondrial respiration suggests that there
may be instances in which NO production, mitochondrial dysfunction, and pathology could be intimately related
[38]. This may depend on the biochemical actions of NO on mitochondria, their signaling consequences, and their
possible relationship to cellular homeostasis and pathophysiology.
Cytochrome C oxidase is situated on the inner membrane of the mitochondrion, where it catalyzes the oxidation of
cytochrome C and the reduction of O2 to water in a process linked to the pumping of protons out of the
mitochondrial matrix. The enzyme contains 2 heme (a and a3) and 2 copper centers (CuA and CuB), of which the
heme iron of cytochrome a3 together with CuB, in their reduced form, form the O2-binding site. NO closely
resembles O2 and therefore can also bind to this site. In the mid 1990s it was demonstrated that NO inhibits the
activity of cytochrome C oxidase [39-41]. This inhibitory effect was shown to be reversible, in competition with O2,
and to occur at concentrations of NO likely to be present physiologically. Thus for example at 30 µM O2
(approximately the tissue concentration of O2) the IC50 of NO for cytochrome C oxidase is 60 nM, whereas at 10
µM (a possible intracellular concentration of O2) [42] the IC50 of NO for the enzyme would be predicted to be
approximately 20 nM. In addition, it has recently been reported [43] that the Ki of NO for the O2-binding site of
cytochrome C oxidase is 0.2 nM, confirming that concentrations of NO that have been detected in tissues (10 to 450
nM) [44, 45] would be sufficient to compete with intracellular O2. The potential biological relevance of the NO-
cytochrome C oxidase interaction has been further highlighted by a number of studies demonstrating inhibition of
respiration by endogenously-generated NO, or its enhancement by inhibitors of NOS in a number of cells, isolated
tissues, and whole animals [46-49]. In studies with vascular endothelial cells in culture it was found that endogenous
concentrations of NO modulate cell respiration in an oxygen-dependent manner [48]. Furthermore, treatment with
the neuropeptide bradykinin, which activates the endothelial isoform of NO synthase (eNOS), generated
concentrations of NO that inhibited respiration further. Conversely, treatment with an inhibitor of NOS resulted in
an immediate increase in O2 consumption, suggesting that endogenous NO interacts with cytochrome C oxidase and
modulates O2 consumption under basal and stimulated conditions. Consistent with studies with isolated cytochrome
C oxidase [50, 51], further work using intact cells suggests that NO interacts with the enzyme in two ways [52]. In
the first case, which occurs at high O2 and low electron turnover in the enzyme, NO interacts primarily with the
prevailing oxidized species of the catalytic cytochrome C oxidase cycle, resulting in an increase in the reduced
fraction of cytochromes cc1 and consequently a rise in the reductive pressure on the NO-free fraction of the enzyme.
This situation, in turn, causes an increase in the electron turnover of the uninhibited fraction of the enzyme, thus
allowing for steady state respiration to be maintained. The second case takes place at low O2, and possibly also at
high O2, if NO levels rise above the physiological nM range. Under these conditions, which favor a high electron
turnover, the high affinity interaction of NO with the reduced species of the catalytic cycle will result in inhibition of
respiration.
3.4 Interaction Between Nitric Oxide and Cytochrome C Oxidase: Generation of
Reactive Oxygen Species by Mitochondria
Experimental evidence accumulated between the late 1960s and late 1970s suggests that a small percentage of the
O2 used by mitochondria is not completely reduced to water but is converted to superoxide anion O2
-• because of the
escape of electrons at complexes I and III of the electron transport chain. Theoretical considerations and
experimental evidence indicate that the redox state of the mitochondrial respiratory chain may be a major
determinant in the control of this process (reviewed by Turrens [53]). Studies using carbon monoxide have also
suggested that the reduction of the electron transport chain as a consequence of cytochrome C oxidase inhibition
may enhance O2
-• formation. [54]. Studies in isolated mitochondria have indicated that treatment with NO generates
O2
-• in a similar manner [55]. It has been suggested that NO acts as a rheostat that sets the concentration of O2 at
which an early reduction of the electron transport chain will occur without inhibition of respiration. When RAW
246.7 cells and HUVECs are incubated at 3% O2 (giving 30uM O2 in the culture medium), the early reduction of the
electron transport chain correlates with an NO-dependent increase in O2
-• levels [52]. These findings suggest that
NO plays a dual role in mitochondrial bioenergetics, on one hand affecting O2 consumption and, on the other,
favoring the generation of O2
-• by decreasing electron flux through the cytochrome C oxidase. Regardless of the
precise mechanism, in the presence of superoxide dismutase (SOD) the NO-induced O2
-•could lead to the formation
of hydrogen peroxide (H2O2) and thus initiate downstream signaling events. In this sense O2
-• generated by the
action of NO on the electron transport chain may represent a second messenger by which mitochondria may
Proc. of SPIE Vol. 6846 684602-8
modulate signal transduction cascades and gene transcription. A similar second messenger role has been ascribed to
O2
-• formed by the action of other cellular oxidases, particularly by NADPH oxidases (reviewed by Griendling et al
[56]). However, the relative contribution to cellular signaling of these sources of O2
-• vis a vis that generated in the
electron transport chain has yet to be assessed.
3.5 Modulation of Mitochondrial Membrane Potential
∆Ψ
m.
Studies in intact lymphoid cells [33] and astrocytes [34]showed that inhibition of respiration by NO results in a
temporary small increase in ∆Ψm. This phenomenon depends on the capacity of some cell types to maintain ATP
levels by glycolysis when respiration is compromised. Generation of a ∆Ψm under these conditions requires entry of
the glycolytically-generated ATP to the mitochondrial matrix via the adenine nucleotide translocator and its
subsequent hydrolysis by the F0F1 ATPase which, now acting in reverse, extrudes protons from the mitochondrial
matrix. An increase in ∆Ψm has previously been detected in association with the initiation of apoptosis [57]. The
possibility that NO may also be involved in this phenomenon is underscored by findings showing that several
proapoptotic factors stimulate NO production [58, 59]. Furthermore, the possibility that a high ∆Ψm promotes the
formation of O2
-• by complex III [60], suggests that this force may also contribute to the NO-stimulated increase in
O2
-• release observed at decreasing O2. Conversely, there are many reports indicating that NO causes mitochondrial
membrane depolarization in association with the induction of apoptosis. Although some of these seemingly
contradictory observations may be attributed to the methodology used to detect changes in ∆Ψm, there may be cases
in which these opposing actions of NO may result from differences in the metabolic or redox environment of the
target cell.
3.6 Activation of AMP-Kinase
Insufficient energy output results in bioenergetic crisis. This phenomenon may stem from a variety of biological
situations, including increased energy demand, restriction of nutrient or oxygen supply (ischemia and hypoxia), and
mitochondrial dysfunction. Bioenergetic crisis causes an increase in intracellular AMP levels, and this in turn leads
to the activation of the AMP-activated protein kinase (AMPK), an enzyme which plays a central role in the control
of intracellular energy metabolism [61]. AMP binding to the enzyme promotes its phosphorylation by the tumor
suppressor LKB1, resulting in full activation. Once activated, the enzyme turns off biosynthetic pathways and at the
same time turns on catabolic pathways, thus conserving ATP levels.
3.7. Nitric oxide and LLLT.
It has been proposed that LLLT might work by photodissociating NO from the cytochrome c oxidase, thereby
reversing the signaling consequences of excessive NO binding [62]. Light can indeed reverse the inhibition caused
by NO binding to cytochrome oxidase, both in isolated mitochondria and in whole cells [63]. Light can also protect
cells against NO-induced cell death. These experiments used light in the visible spectrum, with wavelengths from
600 to 630 nm. NIR also seems to have effects on cytochrome oxidase in conditions where NO is unlikely to be
present.
Light mediated vasodilation was first described in 1968 by R F Furchgott, in his nitric oxide research that lead to his
receipt of a Nobel Prize thirty years later in 1998 [64]. Later studies conducted by other researchers confirmed and
extended Furchgott’s early work and demonstrate the ability of light to influence the localized production or release
of NO and stimulate vasodilation through the effect NO on cGMP. This finding suggests that properly designed
illumination devices may be effective, noninvasive therapeutic agents for patients who would benefit from increased
localized NO availability. However the wavelengths that are most effective on this light mediated release of NO are
different from those used in LLT being in the UVA and blue range [65].
Some wavelengths of light are absorbed by hemoglobin and that illumination can release the NO from hemoglobin
(specifically from the nitrosothiols in the beta chain of the hemoglobin molecule) in red blood cells (RBCs) [66-68]
Since RBCs are continuously delivered to the area of treatment, there is a natural supply of NO that can be released
from each new RBC that passes under the light source and is exposed to the appropriate wavelength of photo
energy. Since the half life of the NO released under the area of illumination is only 2 to 3 seconds, NO release is
very local, preventing the effect of increased NO from being manifested in other portions of the body. Vasodilation
from NO is based its effect on the enzyme guanylate cyclase (GC), which forms cGMP to phosphorylate myosin and
Proc. of SPIE Vol. 6846 684602-9
relax smooth muscle cells in the vascular system. Once available levels of GC are saturated with NO, or once
maximum levels of cGMP are achieved, further vasodilation through illumination will not occur until these biologic
compounds return to their pre-illumination status. Again the wavelengths that have been shown to mediate this
effect tend to be in the UVA and blue ranges not the red and NIR wavelength ranges that are mainly used for LLLT
[69].
Tiina Karu provided experimental evidence [62] that NO was involved in the mechanism of the cellular response to
LLLT in the red region of the spectrum. A suspension of HeLa cells was irradiated with 600-860 nm, or with a diode
laser 820 nm and the number of cells attached to a glass matrix was counted after 30 minute incubation. The NO
donors sodium nitroprusside (SNP), glyceryl trinitrate (GTN), or sodium nitrite (NaNO2) were added to the cellular
suspension before or after irradiation. Treating the cellular suspension with SNP before irradiation significantly
modifies the action spectrum for the enhancement of the cell attachment property and eliminates the light-induced
increase in the number of cells attached to the glass matrix, supposedly by way of binding NO to cytochrome c
oxidase. Other in vivo studies on use of 780-nm for stimulating bone healing in rats [70], the use of 804-nm laser to
decrease damage inflicted in rat hearts after creation of heart attacks [71], have shown significant increases of nitric
oxide in illuminated tissues after LLLT. On the other hand studies have been reported on the use of red and NIR
LLLT to treat mice with arthritis caused by intra-articular injection of zymosan [72], and studies with 660-nm laser
for strokes created in rats [73] have both shown reduction of NO in the tissues. These authors explained this
observation by proposing that LLLT inhibited iNOS.
4. CONSEQUENCES FOR LLLT
Many published papers describe increased blood flow during and after LLLT treatments both in animal models and
in patients. One key question that has not been answered as yet is: does this increased blood flow arise from light
mediated release of NO? If so what is the source? Is it NO that is photodissociated from hemoglobin in circulating
erythrocytes, or NO that is photodissociated from other labile NO stores in the blood vessel wall, or is it derived
from dissociation of NO that has bound to cytochrome c oxidase in the mitochondria of cells in the illuminated area?
It appears that the optimum wavelengths are different for these three processes. Blue light at 441-nm appears to be
best for dissociating NO from hemoglobin, UVA light at 366-nm appears to be best for dissociating NO from blood
vessel walls, and red or NIR light appears to be best for dissociating NO from cytochrome c oxidase.
One observation about the effects of LLLT as it is normally used that needs explanation is the selectivity for injured
or diseased tissues. Illumination of normal tissue in general has little effect. For instance illumination of normal skin
or mucosa does not induce hyperplasia, and illumination of uninjured nerves does not generally induce anesthesia.
This selectivity could be partially explained by the action of light on mitochondria of cells that are injured,
predisposed to apoptosis or hypoxic. In these damaged cells it is possible that the ratio of NO to O2 bound to
cytochrome c oxidase is biased away from O2 and towards NO. If this was the case the mitochondrial respiration
could be reduced dramatically for only small changes in the ration, and dissociation of only small amounts of NO
away from the active sites in the cytochrome c oxidase enzyme would result in large and relatively sustained
increases in respiration and consequent rises in ATP, metabolism and cellular activity.
There are reports that LLLT can induce angiogenesis or growth of new blood vessels that are necessary in wound
healing and especially in repair of ulcers and other non-healing wounds for which LLLY is frequently carried out.
AS mentioned in Section 3.3 the binding of nitric oxide to cytochrome c oxidase can inhibit the stabilization of
hypoxia-inducible factor (HIF) 1-α that would otherwise occur in cells with low oxygen concentrations.
Stabilization of HIF1-α is one of the main mechanism for cells to initiate the formation of new blood vessels as a
response to tissue hypoxia. Vascular endothelial growth factor is an important gene whose transcription is regulated
by HIF1-α.
It has been known for some time that LLLT is particularly effective at reducing swelling or edema in tissues and in
improving lymphatic drainage. This can also be explained by the effect of nitric oxide in activating the lymphatic
drainage and in relaxing the lymphatic smooth muscle cells. Lymphatic endothelial cells (LECs) specifically express
the α1β1 isoform of soluble guanylate cyclase (sGC) [77], and NO induced LEC proliferation, migration, and
cGMP production in LECs are specifically dependent on sGCα1β1. Moreover, the specific sGC inhibitor NS-2028
Proc. of SPIE Vol. 6846 684602-10
completely prevents ultraviolet B-irradiation-induced lymphatic vessel enlargement, edema formation, and skin
inflammation in vivo. These findings identify a crucial role of the NO/sGCα1β1/cGMP pathway in modulating
lymphatic vessel function. Mechanical activity of lymph vessels with or without the endothelium were investigated
with macrophage conditioned medium [78]. Rat peritoneal macrophages stimulated with LPS suppressed
significantly the basal tone of the lymphatic bioassay rings precontracted by U46619. The induced vasodilation of
the lymph nodes was significantly reduced by preincubation of the macrophages with N omega-nitro-L-arginine
methyl ester indomethacin, or dexamethasone. Simultaneous preincubation of L-NAME and indomethacin caused a
synergistic reduction of the M phi-induced vasodilation of the lymphatic bioassay rings. These findings suggest that
macrophages activated by bacterial LPS produce a marked relaxation of lymphatic smooth muscles through the co-
release of nitric oxide and vasodilative prostaglandins.
ACKNOWLEDGEMENTS
M. R. Hamblin was supported by US National Institutes of Health (R01CA/AI838801 and R01 AI050875)
REFERENCES
[1] S. Passarella, E. Casamassima, S. Molinari, D. Pastore, E. Quagliariello, I.M. Catalano and A. Cingolani,
Increase of proton electrochemical potential and ATP synthesis in rat liver mitochondria irradiated in vitro
by helium-neon laser, FEBS Lett 175 (1984) 95-9.
[2] M. Greco, G. Guida, E. Perlino, E. Marra and E. Quagliariello, Increase in RNA and protein synthesis by
mitochondria irradiated with helium-neon laser, Biochem Biophys Res Commun 163 (1989) 1428-34.
[3] D. Pastore, M. Greco, V.A. Petragallo and S. Passarella, Increase in <--H+/e- ratio of the cytochrome c
oxidase reaction in mitochondria irradiated with helium-neon laser, Biochem Mol Biol Int 34 (1994) 817-
26.
[4] W. Yu, J.O. Naim, M. McGowan, K. Ippolito and R.J. Lanzafame, Photomodulation of oxidative
metabolism and electron chain enzymes in rat liver mitochondria, Photochem Photobiol 66 (1997) 866-71.
[5] M.W. Gordon, The correlation between in vivo mitochondrial changes and tryptophan pyrrolase activity,
Arch Biochem Biophys 91 (1960) 75-82.
[6] N.L. Vekshin, Light-dependent ATP synthesis in mitochondria, Biochem Int 25 (1991) 603-11.
[7] S. Passarella, A. Ostuni, A. Atlante and E. Quagliariello, Increase in the ADP/ATP exchange in rat liver
mitochondria irradiated in vitro by helium-neon laser, Biochem Biophys Res Commun 156 (1988) 978-86.
[8] Y. Morimoto, T. Arai, M. Kikuchi, S. Nakajima and H. Nakamura, Effect of low-intensity argon laser
irradiation on mitochondrial respiration, Lasers Surg Med 15 (1994) 191-9.
[9] T. Karu, L. Pyatibrat and G. Kalendo, Irradiation with He-Ne laser increases ATP level in cells cultivated
in vitro, J Photochem Photobiol B 27 (1995) 219-23.
[10] L.E. Bakeeva, V.M. Manteifel, E.B. Rodichev and T.I. Karu, Formation of gigantic mitochondria in human
blood lymphocytes under the effect of an He-Ne laser, Mol Biol (Mosk) 27 (1993) 608-17.
[11] V.M. Manteifel, T.N. Andreichuk, T.I. Karu, P.V. Chelidze and A.V. Zelenin, Activation of transcription
function in lymphocytes after irradiation with He-Ne-laser, Mol Biol (Mosk) 24 (1990) 1067-75.
[12] H. Loevschall and D. Arenholt-Bindslev, Effect of low level diode laser irradiation of human oral mucosa
fibroblasts in vitro, Lasers Surg Med 14 (1994) 347-54.
[13] T.I. Karu and N.I. Afanas'eva, Cytochrome c oxidase as the primary photoacceptor upon laser exposure of
cultured cells to visible and near IR-range light, Dokl Akad Nauk 342 (1995) 693-5.
[14] T.I. Karu and S.F. Kolyakov, Exact action spectra for cellular responses relevant to phototherapy,
Photomed Laser Surg 23 (2005) 355-61.
[15] C.M. Carnevalli, C.P. Soares, R.A. Zangaro, A.L. Pinheiro and N.S. Silva, Laser light prevents apoptosis in
Cho K-1 cell line, J Clin Laser Med Surg 21 (2003) 193-6.
[16] H.T. Whelan, E.V. Buchmann, A. Dhokalia, M.P. Kane, N.T. Whelan, M.T. Wong-Riley, J.T. Eells, L.J.
Gould, R. Hammamieh, R. Das and M. Jett, Effect of NASA light-emitting diode irradiation on molecular
changes for wound healing in diabetic mice, J Clin Laser Med Surg 21 (2003) 67-74.
[17] H.T. Whelan, R.L. Smits, Jr., E.V. Buchman, N.T. Whelan, S.G. Turner, D.A. Margolis, V. Cevenini, H.
Stinson, R. Ignatius, T. Martin, J. Cwiklinski, A.F. Philippi, W.R. Graf, B. Hodgson, L. Gould, M. Kane,
G. Chen and J. Caviness, Effect of NASA light-emitting diode irradiation on wound healing, J Clin Laser
Med Surg 19 (2001) 305-14.
Proc. of SPIE Vol. 6846 684602-11
[18] M.T. Wong-Riley, H.L. Liang, J.T. Eells, B. Chance, M.M. Henry, E. Buchmann, M. Kane and H.T.
Whelan, Photobiomodulation directly benefits primary neurons functionally inactivated by toxins: role of
cytochrome c oxidase, J Biol Chem 280 (2005) 4761-71.
[19] J.T. Eells, M.M. Henry, P. Summerfelt, M.T. Wong-Riley, E.V. Buchmann, M. Kane, N.T. Whelan and
H.T. Whelan, Therapeutic photobiomodulation for methanol-induced retinal toxicity, Proc Natl Acad Sci U
S A 100 (2003) 3439-44.
[20] R.A. Capaldi, F. Malatesta and V.M. Darley-Usmar, Structure of cytochrome c oxidase, Biochim Biophys
Acta 726 (1983) 135-48.
[21] I. Szundi, G.L. Liao and O. Einarsdottir, Near-infrared time-resolved optical absorption studies of the
reaction of fully reduced cytochrome c oxidase with dioxygen, Biochemistry 40 (2001) 2332-9.
[22] D. Pastore, M. Greco and S. Passarella, Specific helium-neon laser sensitivity of the purified cytochrome c
oxidase, Int J Radiat Biol 76 (2000) 863-70.
[23] V.G. Artyukhov, O.V. Basharina, A.A. Pantak and L.S. Sveklo, Effect of helium-neon laser on activity and
optical properties of catalase, Bull Exp Biol Med 129 (2000) 537-40.
[24] S. Passarella, He-Ne laser irradiation of isolated mitochondria, J Photochem Photobiol B 3 (1989) 642-3.
[25] T.N. Raju, The Nobel chronicles. 1998: Robert Francis Furchgott (b 1911), Louis J Ignarro (b 1941), and
Ferid Murad (b 1936), Lancet 356 (2000) 346.
[26] W.M. Xu and L.Z. Liu, Nitric oxide: from a mysterious labile factor to the molecule of the Nobel Prize.
Recent progress in nitric oxide research, Cell Res 8 (1998) 251-8.
[27] S. Moncada and J.D. Erusalimsky, Does nitric oxide modulate mitochondrial energy generation and
apoptosis?, Nat Rev Mol Cell Biol 3 (2002) 214-20.
[28] G.C. Brown, Regulation of mitochondrial respiration by nitric oxide inhibition of cytochrome c oxidase,
Biochim Biophys Acta 1504 (2001) 46-57.
[29] A.J. Kanai, L.L. Pearce, P.R. Clemens, L.A. Birder, M.M. VanBibber, S.Y. Choi, W.C. de Groat and J.
Peterson, Identification of a neuronal nitric oxide synthase in isolated cardiac mitochondria using
electrochemical detection, Proc Natl Acad Sci U S A 98 (2001) 14126-31.
[30] V. Darley-Usmar, The powerhouse takes control of the cell; the role of mitochondria in signal transduction,
Free Radic Biol Med 37 (2004) 753-4.
[31] M.D. Brand, The efficiency and plasticity of mitochondrial energy transduction, Biochem Soc Trans 33
(2005) 897-904.
[32] F. Antunes, A. Boveris and E. Cadenas, On the mechanism and biology of cytochrome oxidase inhibition
by nitric oxide, Proc Natl Acad Sci U S A 101 (2004) 16774-9.
[33] B. Beltran, A. Mathur, M.R. Duchen, J.D. Erusalimsky and S. Moncada, The effect of nitric oxide on cell
respiration: A key to understanding its role in cell survival or death, Proc Natl Acad Sci U S A 97 (2000)
14602-7.
[34] A. Almeida, J. Almeida, J.P. Bolanos and S. Moncada, Different responses of astrocytes and neurons to
nitric oxide: the role of glycolytically generated ATP in astrocyte protection, Proc Natl Acad Sci U S A 98
(2001) 15294-9.
[35] W. Xu, L. Liu, I.G. Charles and S. Moncada, Nitric oxide induces coupling of mitochondrial signalling
with the endoplasmic reticulum stress response, Nat Cell Biol 6 (2004) 1129-34.
[36] A. Almeida, S. Moncada and J.P. Bolanos, Nitric oxide switches on glycolysis through the AMP protein
kinase and 6-phosphofructo-2-kinase pathway, Nat Cell Biol 6 (2004) 45-51.
[37] T. Hagen, C.T. Taylor, F. Lam and S. Moncada, Redistribution of intracellular oxygen in hypoxia by nitric
oxide: effect on HIF1alpha, Science 302 (2003) 1975-8.
[38] D. Brealey, M. Brand, I. Hargreaves, S. Heales, J. Land, R. Smolenski, N.A. Davies, C.E. Cooper and M.
Singer, Association between mitochondrial dysfunction and severity and outcome of septic shock, Lancet
360 (2002) 219-23.
[39] M.W. Cleeter, J.M. Cooper, V.M. Darley-Usmar, S. Moncada and A.H. Schapira, Reversible inhibition of
cytochrome c oxidase, the terminal enzyme of the mitochondrial respiratory chain, by nitric oxide.
Implications for neurodegenerative diseases, FEBS Lett 345 (1994) 50-4.
[40] G.C. Brown and C.E. Cooper, Nanomolar concentrations of nitric oxide reversibly inhibit synaptosomal
respiration by competing with oxygen at cytochrome oxidase, FEBS Lett 356 (1994) 295-8.
[41] M. Schweizer and C. Richter, Nitric oxide potently and reversibly deenergizes mitochondria at low oxygen
tension, Biochem Biophys Res Commun 204 (1994) 169-75.
Proc. of SPIE Vol. 6846 684602-12
[42] M. Tamura, O. Hazeki, S. Nioka and B. Chance, In vivo study of tissue oxygen metabolism using optical
and nuclear magnetic resonance spectroscopies, Annu Rev Physiol 51 (1989) 813-34.
[43] M.G. Mason, P. Nicholls, M.T. Wilson and C.E. Cooper, Nitric oxide inhibition of respiration involves
both competitive (heme) and noncompetitive (copper) binding to cytochrome c oxidase, Proc Natl Acad Sci
U S A 103 (2006) 708-13.
[44] K. Shibuki and D. Okada, Endogenous nitric oxide release required for long-term synaptic depression in
the cerebellum, Nature 349 (1991) 326-8.
[45] T. Malinski, Z. Taha, S. Grunfeld, S. Patton, M. Kapturczak and P. Tomboulian, Diffusion of nitric oxide
in the aorta wall monitored in situ by porphyrinic microsensors, Biochem Biophys Res Commun 193
(1993) 1076-82.
[46] W. Shen, T.H. Hintze and M.S. Wolin, Nitric oxide. An important signaling mechanism between vascular
endothelium and parenchymal cells in the regulation of oxygen consumption, Circulation 92 (1995) 3505-
12.
[47] P.R. Miles, L. Bowman and L. Huffman, Nitric oxide alters metabolism in isolated alveolar type II cells,
Am J Physiol 271 (1996) L23-30.
[48] E. Clementi, G.C. Brown, N. Foxwell and S. Moncada, On the mechanism by which vascular endothelial
cells regulate their oxygen consumption, Proc Natl Acad Sci U S A 96 (1999) 1559-62.
[49] K.E. Loke, P.I. McConnell, J.M. Tuzman, E.G. Shesely, C.J. Smith, C.J. Stackpole, C.I. Thompson, G.
Kaley, M.S. Wolin and T.H. Hintze, Endogenous endothelial nitric oxide synthase-derived nitric oxide is a
physiological regulator of myocardial oxygen consumption, Circ Res 84 (1999) 840-5.
[50] P. Sarti, A. Giuffre, E. Forte, D. Mastronicola, M.C. Barone and M. Brunori, Nitric oxide and cytochrome c
oxidase: mechanisms of inhibition and NO degradation, Biochem Biophys Res Commun 274 (2000) 183-7.
[51] A. Giuffre, M.C. Barone, D. Mastronicola, E. D'Itri, P. Sarti and M. Brunori, Reaction of nitric oxide with
the turnover intermediates of cytochrome c oxidase: reaction pathway and functional effects, Biochemistry
39 (2000) 15446-53.
[52] M. Palacios-Callender, M. Quintero, V.S. Hollis, R.J. Springett and S. Moncada, Endogenous NO regulates
superoxide production at low oxygen concentrations by modifying the redox state of cytochrome c oxidase,
Proc Natl Acad Sci U S A 101 (2004) 7630-5.
[53] J.F. Turrens, Mitochondrial formation of reactive oxygen species, J Physiol 552 (2003) 335-44.
[54] J. Zhang and C.A. Piantadosi, Mitochondrial oxidative stress after carbon monoxide hypoxia in the rat
brain, J Clin Invest 90 (1992) 1193-9.
[55] J.J. Poderoso, M.C. Carreras, C. Lisdero, N. Riobo, F. Schopfer and A. Boveris, Nitric oxide inhibits
electron transfer and increases superoxide radical production in rat heart mitochondria and
submitochondrial particles, Arch Biochem Biophys 328 (1996) 85-92.
[56] K.K. Griendling, D. Sorescu, B. Lassegue and M. Ushio-Fukai, Modulation of protein kinase activity and
gene expression by reactive oxygen species and their role in vascular physiology and pathophysiology,
Arterioscler Thromb Vasc Biol 20 (2000) 2175-83.
[57] K. Banki, E. Hutter, N.J. Gonchoroff and A. Perl, Elevation of mitochondrial transmembrane potential and
reactive oxygen intermediate levels are early events and occur independently from activation of caspases in
Fas signaling, J Immunol 162 (1999) 1466-79.
[58] M.S. Williams, S. Noguchi, P.A. Henkart and Y. Osawa, Nitric oxide synthase plays a signaling role in
TCR-triggered apoptotic death, J Immunol 161 (1998) 6526-31.
[59] P. Secchiero, A. Gonelli, C. Celeghini, P. Mirandola, L. Guidotti, G. Visani, S. Capitani and G. Zauli,
Activation of the nitric oxide synthase pathway represents a key component of tumor necrosis factor-
related apoptosis-inducing ligand-mediated cytotoxicity on hematologic malignancies, Blood 98 (2001)
2220-8.
[60] A. Galkin, A. Higgs and S. Moncada, Nitric oxide and hypoxia, Essays Biochem 43 (2007) 29-42.
[61] D. Carling, The AMP-activated protein kinase cascade--a unifying system for energy control, Trends
Biochem Sci 29 (2004) 18-24.
[62] T.I. Karu, L.V. Pyatibrat and N.I. Afanasyeva, Cellular effects of low power laser therapy can be mediated
by nitric oxide, Lasers Surg Med 36 (2005) 307-14.
[63] V. Borutaite, A. Budriunaite and G.C. Brown, Reversal of nitric oxide-, peroxynitrite- and S-nitrosothiol-
induced inhibition of mitochondrial respiration or complex I activity by light and thiols, Biochim Biophys
Acta 1459 (2000) 405-12.
Proc. of SPIE Vol. 6846 684602-13
[64] S.J. Ehrreich and R.F. Furchgott, Relaxation of mammalian smooth muscles by visible and ultraviolet
radiation, Nature 218 (1968) 682-4.
[65] H. Chaudhry, M. Lynch, K. Schomacker, R. Birngruber, K. Gregory and I. Kochevar, Relaxation of
vascular smooth muscle induced by low-power laser radiation, Photochem Photobiol 58 (1993) 661-9.
[66] R. Mittermayr, A. Osipov, C. Piskernik, S. Haindl, P. Dungel, C. Weber, Y.A. Vladimirov, H. Redl and
A.V. Kozlov, Blue laser light increases perfusion of a skin flap via release of nitric oxide from hemoglobin,
Mol Med 13 (2007) 22-9.
[67] L. Vladimirov, A., G.I. Klebanov, G.G. Borisenko and A.N. Osipov, Molecular and cellular mechanisms of
the low intensity laser radiation effect, Biofizika 49 (2004) 339-50.
[68] Y.A. Vladimirov, A.N. Osipov and G.I. Klebanov, Photobiological principles of therapeutic applications of
laser radiation, Biochemistry (Mosc) 69 (2004) 81-90.
[69] G.G. Borisenko, A.N. Osipov, K.D. Kazarinov and A. Vladimirov Yu, Photochemical reactions of nitrosyl
hemoglobin during exposure to low-power laser irradiation, Biochemistry (Mosc) 62 (1997) 661-6.
[70] G.A. Guzzardella, M. Fini, P. Torricelli, G. Giavaresi and R. Giardino, Laser stimulation on bone defect
healing: an in vitro study, Lasers Med Sci 17 (2002) 216-20.
[71] H. Tuby, L. Maltz and U. Oron, Modulations of VEGF and iNOS in the rat heart by low level laser therapy
are associated with cardioprotection and enhanced angiogenesis, Lasers Surg Med 38 (2006) 682-8.
[72] Y. Moriyama, E.H. Moriyama, K. Blackmore, M.K. Akens and L. Lilge, In vivo study of the inflammatory
modulating effects of low-level laser therapy on iNOS expression using bioluminescence imaging,
Photochem Photobiol 81 (2005) 1351-5.
[73] M.C. Leung, S.C. Lo, F.K. Siu and K.F. So, Treatment of experimentally induced transient cerebral
ischemia with low energy laser inhibits nitric oxide synthase activity and up-regulates the expression of
transforming growth factor-beta 1, Lasers Surg Med 31 (2002) 283-8.
[74] Y. Moriyama, E.H. Moriyama, K. Blackmore, M.K. Akens and L. Lilge, In vivo Study of the Inflammatory
Modulating Effects of Low Level Laser Therapy on iNOS Expression Using Bioluminescence Imaging,
Photochem Photobiol (2005).
[75] T.I. Karu, L.V. Pyatibrat and G.S. Kalendo, Donors of NO and pulsed radiation at lambda = 820 nm exert
effects on cell attachment to extracellular matrices, Toxicol Lett 121 (2001) 57-61.
[76] C.F. Rizzi, J.L. Mauriz, D.S. Freitas Correa, A.J. Moreira, C.G. Zettler, L.I. Filippin, N.P. Marroni and J.
Gonzalez-Gallego, Effects of low-level laser therapy (LLLT) on the nuclear factor (NF)-kappaB signaling
pathway in traumatized muscle, Lasers Surg Med 38 (2006) 704-13.
[77] K. Kajiya, R. Huggenberger, I. Drinnenberg, B. Ma and M. Detmar, Nitric oxide mediates lymphatic vessel
activation via soluble guanylate cyclase {alpha}1{beta}1-impact on inflammation, Faseb J (2007).
[78] H. Wang, Activated macrophage-mediated endogenous prostaglandin and nitric oxide-dependent relaxation
of lymphatic smooth muscles, Jpn J Physiol 47 (1997) 93-100.
Proc. of SPIE Vol. 6846 684602-14
... NO is a potent endogenous vasodilator involved in the effects of R/NIR light on the vascular system. R/NIR light exposure probably causes vasodilatory effects by photo-dissociating NO from CCO, the NO stores in endothelial cells in blood vessels, and/or the hemoglobin in circulating erythrocytes 13 . Red and NIR light increases vasodilation in healthy and diabetic mice by releasing NO-bound substances (S-nitrosothiols or non-heme iron nitrosyl complexes) from the vascular endothelium 14,15 . ...
... Several studies have demonstrated that R/NIR-induced increase in local blood flow is either associated with localized NO release or stimulation of NO synthase gene expression in cell systems 11,13,14,35 . R/NIR light has been shown to protect cardiomyocytes from hypoxia and reoxygenation damage using a NO-dependent mechanism 18 , cause vasodilation, and restore endothelial dysfunction in blood vessels of healthy and diabetic mice by releasing NO-bound substances from the vascular endothelium. ...
... Quantifying NO or its metabolites in blood/plasma samples would be needed to validate these hypotheses. Although the mechanisms behind increased blood flow following PBM are not entirely understood, our study indicates that NIR exposure increases the local skin blood flow, which might be NO-dependent 8,13 . Our observations that both intermittent and continuous modes of irradiation increase blood flow are in line with the results in a recent study by Gavish et al. 23 , which reported a 54 % increase in local skin blood flow in the palm after 5 minutes of wrist irradiation with 830 nm NIR (irradiance: 55 mW /cm 2 ; fluence: 16.5 J/cm 2 ). ...
Preprint
Full-text available
Photobiomodulation causes an immediate increase in local blood flow. This study aimed to investigate the effect of 890 nm NIR exposure on local skin blood flow in young and middle-aged healthy subjects. In this placebo-controlled clinical trial, 12 young and 12 middle-aged healthy subjects received either continuous or intermittent NIR exposure (890 nm, 5.1 mW /cm 2 , 4.6 J/cm 2 , and 35.9 J total energy) on the skin of the upper lateral arm. The continuous exposure experiment, used in young subjects only, applied 30 minutes of continuous NIR light (experiment 1). The intermittent exposure experiment, used in both age groups, applied NIR light through 10 cycles of 3 minutes NIR exposure and 2 minutes OFF (for recording blood flow), resulting in 50 minutes of total time. Laser Doppler flowmetry and thermal images were used to monitor local blood flow and skin temperature. In young subjects, continuous NIR exposure compared to placebo significantly increased blood flow for the first 20 minutes post-exposure. Further, in young and middle-aged subjects, intermittent exposure increased blood flow during the whole exposure period and 15 minutes post-exposure. In young subjects, blood flow after continuous NIR exposure was significantly higher than intermittent NIR exposure only for the first 10 minutes. For intermittent exposure comparison between the two age groups, the blood flow was significantly higher in middle-aged subjects. We conclude that NIR PBM increases local skin blood flow in young and middle-aged subjects. The mode of NIR irradiation and the subjects’ age influenced the local skin blood flow response.
... The first-studied pathway shone a light on CCOs. When the electrons in the metal centers of CCO are excited by photon absorption (Figure 2) [44,45,[47][48][49][50][51][52], nitrous oxide (NO) from CCO's binuclear center (heme a3/CuB) is photodissociated. Decreasing amounts of NO, a known electron transport inhibitor in the ETC, raise the mitochondrial membrane potential (MMP), consequently increasing the proton gradient and ATP production [52]. ...
... When the electrons in the metal centers of CCO are excited by photon absorption (Figure 2) [44,45,[47][48][49][50][51][52], nitrous oxide (NO) from CCO's binuclear center (heme a3/CuB) is photodissociated. Decreasing amounts of NO, a known electron transport inhibitor in the ETC, raise the mitochondrial membrane potential (MMP), consequently increasing the proton gradient and ATP production [52]. This upregulates the function of reactive oxygen species (ROS) and calcium ions as secondary messengers, resulting in the activation of transcription factors and cell proliferation signaling molecules such as nuclear factor kappa-light-chain-enhancer of activated B cells (NF-kB) ( Figure 2) [53,54]. ...
... In this second pathway, light appears to be absorbed by temperature-gated calcium channels, causing an increase in cytosolic calcium but a decrease in mitochondrial calcium [48]. . This is a schematic diagram of commonly known near-infrared light targets and the biological effects of transcranial photobiomodulation in the management of AD pathology and associated disease processes [44,45,[47][48][49][50][51][52]. Abbreviations: ROS, reactive oxygen species; NO, nitric oxide; eNOS, endothelial nitric oxide synthase; SOD, superoxide dismutase; GPx, glutamate peroxidase; CcO, cytochrome C oxidase; TGF-β, transforming growth factor-beta; TNF-α, tumor necrosis factor-alpha; IFN-γ, interferon-gamma. ...
Article
Full-text available
Alzheimer’s disease (AD) is a neurodegenerative disease and the world’s primary cause of dementia, a condition characterized by significant progressive declines in memory and intellectual capacities. While dementia is the main symptom of Alzheimer’s, the disease presents with many other debilitating symptoms, and currently, there is no known treatment exists to stop its irreversible progression or cure the disease. Photobiomodulation has emerged as a very promising treatment for improving brain function, using light in the range from red to the near-infrared spectrum depending on the application, tissue penetration, and density of the target area. The goal of this comprehensive review is to discuss the most recent achievements in and mechanisms of AD pathogenesis with respect to neurodegeneration. It also provides an overview of the mechanisms of photobiomodulation associated with AD pathology and the benefits of transcranial near-infrared light treatment as a potential therapeutic solution. This review also discusses the older reports and hypotheses associated with the development of AD, as well as some other approved AD drugs.
... It has been confirmed to speed up collateral circulation and promote microcirculation. [18] demonstrated and proved the ability of light to have a great impact on the localized release of nitric oxide (NO) and lead to vasodilation through the effect of NO on cyclic guanosine monophosphate cGMP. ...
... They concluded that laser irradiation could cause marked dilatation in vessels and expansion in the bloodstream of arterioles through a direct effect on VSMC as it reduced [Ca2+] ions in microvascular smooth muscles. A systemic review conducted byHamblin (2008) ...
Article
Aims of the study. This study aimed to studying the effect of Low level laser therapy (LLLT) on vertebral artery blood flow in elderly with cervical disc degeneration. The effect of degenerative changes in the cervical spine on the vertebral arteries and on the supply of blood to the brain stem and the inner ear is now well known. In 20–30% of cases, ischemia of the posterior cranial fossa caused by vertebrobasilar insufficiency leads to death. Materials and methods. Research involved sixty patients of both gender, whose ages between 60 and 75 years, how are suffering from cervical disc degeneration. Subjects were randomly assigned to two equal groups. Group (A), Study Group, received low level laser therapy (LLLT) with a wavelength of 830 nm and power of 200 mW on vertebral artery bilaterally and routine physical therapy exercises. Group (B), the control group, received only routine physical therapy exercises. Both groups received three sessions per week for 12 weeks. Blood flow in both vertebral arteries was measured by assessing Resistivity Index using ultrasound Doppler, pre and post treatment for both groups. Results. There was no significant difference between groups pretreatment (p > 0.05). There was a significant decrease in resistivity index of right and left vertebral arteries of group A compared with that of group B post treatment (p < 0.01). There was a significant decrease in resistivity index of right and left vertebral arteries of group A and B post treatment compared with pre treatment (p < 0.001). The percent of change of resistivity index of right and left vertebral arteries in group A was 8.7 and 6.12% respectively and that in group B was 2.63 and 1.45% respectively. Conclusion. When LLLT was applied using the study's settings, older patients with cervical disc degeneration experienced bilateral improvements in blood flow through the vertebral arteries.
... Furthermore, NO produced by COX has the ability to increase ATP generation by raising mitochondrial membrane potential and oxygen consumption [350][351][352][353]. as well as eliciting a natural hemodynamic response to increased oxygen supply to the human brain [353]. However, PBM effects may be mediated by pathways other than COX under specific situations, as evidenced by the wide metabolomics effects of PBM on the rat brain [355]. ...
Article
This narrative review summerizes the mitochondria dyafunction related to neurodegenerative disorders and summerizes the different treatment methods of them
Article
Full-text available
Photobiomodulation (PBM) has been applied in biomedical technology to improve cellular responses. Specifically in sports medicine, it is used to accelerate metabolic and structural repair and adaptation in skeletal muscle under stress overload. Currently, PBM has been associated with static Magnetic Field (sMF) in clinical applications, enhancing the effects displayed by PBM when used in isolation. However, the biochemical and molecular effects of PBM-sMF in myoblasts remain unknown. This study aimed to investigate the effects of PBM combined with static magnetic field (PBM-sMF) at different doses in C2C12 muscle cells in the presence or absence of hydrogen peroxide (H2O2), a standard oxidant. Different spectrophotometric and fluorometric assays were conducted after cellular treatments. PBM-sMF was shown to be effective compared to H2O2 regarding cell viability and release of nitric oxide (NO), dsDNA, and reactive oxygen species (ROS) levels. It positively modulated mitochondrial membrane potential (ΔΨm) and cytochrome c oxidase (COX) activity under normal conditions and restored both to normal levels when impacted by H2O2. Regarding apoptosis, the recovery in viable cells observed on PBM-sMF treated cells was dose-dependent. In conclusion, PBM-sMF has a biphasic effect in normal and oxidative environments and may differently modulate myoblast cells depending on their redox status.
Chapter
This chapter explores the range of benefits that relate to laser-assisted oral hard tissue management and details aspects of each wavelength in delivering adjunctive therapy. Of the currently available wavelengths of dental lasers, at the moment, only three can be used for hard tissue.
Article
Mitochondria play a vital role in the nervous system, as they are responsible for generating energy in the form of ATP and regulating cellular processes such as calcium (Ca2+) signaling and apoptosis. However, mitochondrial dysfunction can lead to oxidative stress (OS), inflammation, and cell death, which have been implicated in the pathogenesis of various neurological disorders. In this article, we review the main functions of mitochondria in the nervous system and explore the mechanisms related to mitochondrial dysfunction. We discuss the role of mitochondrial dysfunction in the development and progression of some neurological disorders including Parkinson's disease (PD), multiple sclerosis (MS), Alzheimer's disease (AD), depression, and epilepsy. Finally, we provide an overview of various current treatment strategies that target mitochondrial dysfunction, including pharmacological treatments, phototherapy, gene therapy, and mitotherapy. This review emphasizes the importance of understanding the role of mitochondria in the nervous system and highlights the potential for mitochondrial-targeted therapies in the treatment of neurological disorders. Furthermore, it highlights some limitations and challenges encountered by the current therapeutic strategies and puts them in future perspective.
Article
Full-text available
Rheumatoid arthritis (RA) and osteoarthritis (OA) have a significant impact on the quality of life of patients around the world, causing significant pain and disability. Furthermore, the drugs used to treat these conditions frequently have side effects that add to the patient's burden. Photobiomodulation (PBM) has emerged as a promising treatment approach in recent years. PBM effectively reduces inflammation by utilizing near-infrared light emitted by lasers or LEDs. In contrast to photothermal effects, PBM causes a photobiological response in cells, which regulates their functional response to light and reduces inflammation. PBM's anti-inflammatory properties and beneficial effects in arthritis treatment have been reported in numerous studies, including animal experiments and clinical trials. PBM's effectiveness in arthritis treatment has been extensively researched in arthritis-specific cells. Despite the positive results of PBM treatment, questions about specific parameters such as wavelength, dose, power density, irradiation time, and treatment site remain. The goal of this comprehensive review is to systematically summarize the mechanisms of PBM in arthritis treatment, the development of animal arthritis models, and the anti-inflammatory and joint function recovery effects seen in these models. The review also goes over the evaluation methods used in clinical trials. Overall, this review provides valuable insights for researchers investigating PBM treatment for arthritis, providing important references for parameters, model techniques, and evaluation methods in future studies.
Article
Full-text available
Since it was first realized that biological energy transduction involves oxygen and ATP, opinions about the amount of ATP made per oxygen consumed have continually evolved. The coupling efficiency is crucial because it constrains mechanistic models of the electron-transport chain and ATP synthase, and underpins the physiology and ecology of how organisms prosper in a thermodynamically hostile environment. Mechanistically, we have a good model of proton pumping by complex III of the electron-transport chain and a reasonable understanding of complex IV and the ATP synthase, but remain ignorant about complex I. Energy transduction is plastic: coupling efficiency can vary. Whether this occurs physiologically by molecular slipping in the proton pumps remains controversial. However, the membrane clearly leaks protons, decreasing the energy funnelled into ATP synthesis. Up to 20% of the basal metabolic rate may be used to drive this basal leak. In addition, UCP1 (uncoupling protein 1) is used in specialized tissues to uncouple oxidative phosphorylation, causing adaptive thermogenesis. Other UCPs can also uncouple, but are tightly regulated; they may function to decrease coupling efficiency and so attenuate mitochondrial radical production. UCPs may also integrate inputs from different fuels in pancreatic β-cells and modulate insulin secretion. They are exciting potential targets for treatment of obesity, cachexia, aging and diabetes.
Article
Full-text available
The mitochondrion is a key organelle in the control of cell death. Nitric oxide (NO) inhibits complex IV in the respiratory chain and is reported to possess both proapoptotic and antiapoptotic actions. We investigated the effects of continuous inhibition of respiration by NO on mitochondrial energy status and cell viability. Serum-deprived human T cell leukemia (Jurkat) cells were exposed to NO at a concentration that caused continuous and complete (∼85%) inhibition of respiration. Serum deprivation caused progressive loss of mitochondrial membrane potential (Δψm) and apoptotic cell death. In the presence of NO, Δψm was maintained compared to controls, and cells were protected from apoptosis. Similar results were obtained by using staurosporin as the apoptotic stimulus. As exposure of serum-deprived cells to NO progressed (>5 h), however, Δψm fell, correlating with the appearance of early apoptotic features and a decrease in cell viability. Glucose deprivation or iodoacetate treatment of cells in the presence of NO resulted in a collapse of Δψm, demonstrating involvement of glycolytic ATP in its maintenance. Under these conditions cell viability also was decreased. Treatment with oligomycin and/or bongkrekic acid indicated that the maintenance of Δψm during exposure to NO is caused by reversal of the ATP synthase and other electrogenic pumps. Thus, blockade of complex IV by NO initiates a protective action in the mitochondrion to maintain Δψm; this results in prevention of apoptosis. It is likely that during cellular stress involving increased generation of NO this compound will trigger a similar sequence of events, depending on its concentration and duration of release.
Article
Full-text available
To better understand the mechanisms of tissue injury during and after carbon monoxide (CO) hypoxia, we studied the generation of partially reduced oxygen species (PROS) in the brains of rats subjected to 1% CO for 30 min, and then reoxygenated on air for 0-180 min. By determining H2O2-dependent inactivation of catalase in the presence of 3-amino-1,2,4-triazole (ATZ), we found increased H2O2 production in the forebrain after reoxygenation. The localization of catalase to brain microperoxisomes indicated an intracellular site of H2O2 production; subsequent studies of forebrain mitochondria isolated during and after CO hypoxia implicated nearby mitochondria as the source of H2O2. In the mitochondria, two periods of PROS production were indicated by decreases in the ratio of reduced to oxidized glutathione (GSH/GSSG). These periods of oxidative stress occurred immediately after CO exposure and 120 min after reoxygenation, as indicated by 50 and 43% decreases in GSH/GSSG, respectively. The glutathione depletion data were supported by studies of hydroxyl radical generation using a salicylate probe. The salicylate hydroxylation products, 2,3 and 2,5-dihydroxybenzoic acid (DHBA), were detected in mitochondria from CO exposed rats in significantly increased amounts during the same time intervals as decreases in GSH/GSSG. The DHBA products were increased 3.4-fold immediately after CO exposure, and threefold after 120 min reoxygenation. Because these indications of oxidative stress were not prominent in the postmitochondrial fraction, we propose that PROS generated in the brain after CO hypoxia originate primarily from mitochondria. These PROS may contribute to CO-mediated neuronal damage during reoxygenation after severe CO intoxication.
Article
Alveolar type II cells may be exposed to nitric oxide (-NO) from external sources, and these cells can also generate -NO. Therefore we studied the effects of altering -NO levels on various type II cell metabolic processes. Incubation of cells with the -NO generator, S-nitroso-AT-acetylpenicillamine (SNAP; 1 mM), leads to reductions of 60-70% in the synthesis of disaturated phosphatidylcholines (DSPC) and cell ATP levels. Cellular oxygen consumption, an indirect measure of cell ATP synthesis, is also reduced by SNAP. There is no direct effect of SNAP on lung mitochondrial ATP synthesis, suggesting that -NO does not directly inhibit this process. On the other hand, incubation of cells with NG-nitro-L-arginine methyl ester (L-NAME), an inhibitor of nitric oxide synthase (NOS), the enzyme responsible for -NO synthesis, results in increases in DSPC synthesis, cell ATP content, and cellular oxygen consumption. The L-NAME effects are reversed by addition of L-arginine, the substrate for NOS. Production of NO by type II cells is inhibited by L-NAME, a better inhibitor of constitutive NOS (cNOS) than inducible NOS (iNOS), and is reduced in the absence of external calcium. Aminoguanidine, a specific inhibitor of iNOS, has no effect on cell ATP content or on -NO production. These results indicate that alveolar type II cell lipid and energy metabolism can be affected by -NO and suggest that there may be cNOS activity in these cells.
Article
Nitric oxide (·NO) released byS-nitrosoglutathione (GSNO) inhibited enzymatic activities of rat heart mitochondrial membranes. Cytochrome oxidase activity was inhibited to one-half at an effective ·NO concentration of 0.1 μM, while succinate– and NADH–cytochrome-creductase activities were half-maximally inhibited at 0.3 μM·NO. Submitochondrial particles treated with ·NO (either from GSNO or from a pure solution) showed increased[formula]and H2O2production when supplemented with succinate alone, at rates that were comparable to those of control particles with added succinate and antimycin. Rat heart mitochondria treated with ·NO also showed increased H2O2production. Cytochrome spectra and decreased enzymatic activities in the presence of ·NO are consistent with a multiple inhibition of mitochondrial electron transfer at cytochrome oxidase and at the ubiquinone–cytochromebregion of the respiratory chain, the latter leading to the increased[formula]production. Electrochemical detection showed that the buildup of a ·NO concentration from GSNO was interrupted by submitochondrial particles supplemented with succinate and antimycin and was restored by addition of superoxide dismutase. The inhibitory effect of ·NO on cytochrome oxidase was also prevented under the same conditions. Apparently, mitochondrial[formula]reacts with ·NO to form peroxynitrite and, by removing ·NO, reactivates the previously inhibited cytochrome oxidase. It is suggested that, at physiological concentrations of ·NO, inhibition of electron transfer, ·NO-induced[formula]production, and ONOO⁻formation participate in the regulatory control of mitochondrial oxygen uptake.
Article
The effects of low level laser (LLL) irradiation on the proliferation of human buccal fibroblasts were studied. A standardized LLL set-up was developed (812 nm, 4.5 ± 0.5 mW/cm2). Cultures in petridishes were divided into eight groups (1 group served as control). On day 6 after seeding, routine growth medium was replaced with PBS for 1/2 hour. At the beginning of this period, LLL irradiation was performed for 0, 1, 3, 10, 32, 100, 316, or 1,000 seconds, respectively—corresponding to the radiant exposures 0, 4.5, 13.5, 45, 144, 450, 1,422, 4,500 mJ/cm2. Subsequently the cells received 3H-dT in fresh medium for 16 hours DNA-incorporation. Scintillations from tritium and total protein concentration per culture dish were determined. The individual 3H-cpm/protein-concentration ratios were calculated in % of control. Three experiments were performed (N = 151). Following LLL exposure the H-cpm/protein ratio was increased with maximum cpm/protein ratio (132.5% ± 10.6% SEM) in the group receiving 450 mJ/cm2 (P < 0.03 nonparametric Kruskal Wallis one-way ANOVA-test). This study demonstrated an increased incorporation on tritiated thymidine in cultured human oral fibroblasts following LLL exposure and suggests that LLL irradiation can induce increased DNA Synthesis. © 1994 Wiley-Liss, Inc.
Article
Light-dependent ATP synthesis was studied in an illuminated suspension of rat liver mitochondria. The action of light was shown to lead to an increase in the ATP content in the absence of oxidisable substrates and in the presence of high (hundreds of microM) ADP concentrations in the medium. At a relatively low (50 microM) ADP concentration, efficient light-dependent phosphorylation was observed in the presence of alpha-ketoglutarate. Prolonged illumination stimulated ATP hydrolysis. Rotenone, antimycin, azide, dicyclohexylcarbodiimide, and oligomycin inhibited the light-dependent phosphorylation almost completely. The level of ATP decreased under the action of 2,4-dinitrophenol in the dark but was restored by high light intensities. Blue light, 436 nm, was most efficient to produce light-dependent phosphorylation. It is assumed that quanta of vibrational excitation formed in the course of vibrational relaxation and the internal conversion of photoexcited flavoproteins and cytochromes are transferred to the ATP-synthetase and "eject" ATP from the active center, thus shifting the enzymatic reaction to ATP production.
Article
Conjunctive stimulation of climbing and parallel fibres in the cerebellum evokes a long-term depression of parallel-fibre Purkinje-cell transmission, a phenomenon implicated as the cellular mechanism for cerebellar motor learning. It is suspected that the increase in cyclic GMP concentration that occurs after activation of climbing fibres is required to evoke long-term depression. Excitatory amino acids are known to cause the release of nitric oxide (NO), resulting in elevation of the cGMP level in the cerebellum. Here we report that endogenous NO is released after stimulation of climbing fibres, that long-term depression evoked by conjunctive stimulation of parallel and climbing fibres is blocked by haemoglobin (which strongly binds NO) or L-NG-monomethyl-arginine (an inhibitor of NO synthase), and that exogenous NO or cGMP can substitute for the stimulation of climbing fibres to cause long-term depression in rat cerebellar slices. These results demonstrate that the release of endogenous NO is essential for the induction of synaptic plasticity in the cerebellum.