Next Article in Journal
Epidemiological Characteristics of 101,471 Patients Hospitalized with Chronic Obstructive Pulmonary Disease (COPD) in Poland in 2019: Multimorbidity, Duration of Hospitalization, In-Hospital Mortality
Previous Article in Journal
The Clinical Significance of Aspergillus Detected in Lower-Respiratory-Tract Samples of Critically Ill COVID-19-Positive Patients
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Constrained Disorder Principle Accounts for the Variability That Characterizes Breathing: A Method for Treating Chronic Respiratory Diseases and Improving Mechanical Ventilation

1
Faculty of Medicine, Hebrew University, Jerusalem P.O. Box 1200, Israel
2
Department of Medicine, Hadassah Medical Center, Jerusalem P.O. Box 1200, Israel
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Adv. Respir. Med. 2023, 91(5), 350-367; https://doi.org/10.3390/arm91050028
Submission received: 21 July 2023 / Revised: 4 September 2023 / Accepted: 5 September 2023 / Published: 9 September 2023

Abstract

:

Highlights

What are the main findings?
  • The constrained disorder principle (CDP) defines systems by their inherent disorder bounded by variable boundaries.
What is the implication of the main finding?
  • The present paper describes the mechanisms of breathing and cellular respiration, focusing on their inherent variability and how the CDP accounts for the variability in breathing and respiration.
  • The article describes using CDP-based artificial intelligence platforms to augment the respiratory process’s efficiency and treat respiratory diseases.

Abstract

Variability characterizes breathing, cellular respiration, and the underlying quantum effects. Variability serves as a mechanism for coping with changing environments; however, this hypothesis does not explain why many of the variable phenomena of respiration manifest randomness. According to the constrained disorder principle (CDP), living organisms are defined by their inherent disorder bounded by variable boundaries. The present paper describes the mechanisms of breathing and cellular respiration, focusing on their inherent variability. It defines how the CDP accounts for the variability and randomness in breathing and respiration. It also provides a scheme for the potential role of respiration variability in the energy balance in biological systems. The paper describes the option of using CDP-based artificial intelligence platforms to augment the respiratory process’s efficiency, correct malfunctions, and treat disorders associated with the respiratory system.

Graphical Abstract

1. Introduction

Variability characterizes the respiration process at the cellular level and in the lungs. It manifests randomly and lacks a regular pattern [1,2]. Quantum effects linked to randomness may underlie some of the mechanisms involved in respiration [3]. Living organisms are defined by the constrained disorder principle (CDP), and complex systems function by their inherent disorder within variable parameters [4].
This paper reviews the variations at different levels of the respiration process and how CDP accounts for them. The paper describes the potential of CDP-based artificial intelligence (AI) systems for correcting diseased states that impact respiration. We selected the relevant studies describing the role of variability in cellular respiration and lung functions and the relevant studies on CDP-based AI’s use to improve organ malfunctions.

1.1. The Constrained Disorder Principle Defines Biological Processes

Variability characterizes the proper function of biological systems. It underlies the function of multiple processes in the genome, microtubules, cellular functions, and whole organs [5,6,7,8,9,10,11,12,13,14,15,16,17,18,19]. The CDP defines living organisms by their degree of disorder bounded by dynamic, continuously changing random boundaries [4]. The principle elucidates the multiple processes that underlie systems’ adaptability, flexibility, and energy use [20]. The CDP can correct systems’ malfunctions because diseased states evolve from low or high degrees of variability. It serves as a platform for second-generation AI systems designed to augment systems’ efficiency [21,22,23].

1.2. The Constrained Disorder Principle Accounts for the Stochasticity in Respiration

Per the CDP, variability, when kept within borders, defines living organisms and is essential for proper respiration [4]. Variability in breathing exists in the resting state and characterizes all organisms, supporting the CDP-based concept that variability is mandatory for proper respiration. It raises the issue of why variability occurs under what seems to be stable and unchanged conditions in the same individual and the benefits of having a degree of variability for the system.
Variability is necessary for assisting with continuous internal and external perturbations in the respiratory systems and is part of the adaptability and flexibility of complex systems [4]. What looks like stable conditions is a sum of numerous continuously changing perturbations mandating a degree of variability for maintaining a proper function. However, one would expect it to take specific patterns of response. The variability seems random and lacks fixed patterns for whole lung function, and cellular respiration implies that it may be part of a CDP-based energy efficiency mechanism of complex systems [20,24]. This mechanism requires dynamic boundaries to the variability, which adapts to the changes in the internal and external environments [4].
In normal subjects, breathing patterns vary in rate, rhythm, depth, and duration [25,26,27]. Mechanical and chemical changes within the respiratory system alter variability [28,29,30]. Breathing control involves complex mechanisms that balance the opposing goals of efficiency, redundancy, responsiveness, and stability [31]. This variability is maintained by a central neural mechanism and feedback loops between lung vagal sensory receptors and arterial chemoreceptors [23]. Respiratory rate, tidal volume, and airflow profiles vary due to inputs, positive and negative feedback loops, internal pacemakers, and non-linear interactions between components [31].
Variability analysis is a method to describe the differences between data groups [32]. Two groups of time-series data can have the same mean values but different variability. Breathing pattern variability can be quantified using quantitative methods, including coefficients of variation and Poincaré plots [31,32,33]. Using the Poincaré plot analysis, breathing patterns are dynamically analyzed breath-by-breath in real time [34].
Breathing-related physiological variables, such as tidal volume (VT) and respiratory rate, exhibit significant breath-to-breath variability [2,35,36]. Cycle-by-cycle variations of respiratory physiological variables are highly reproducible in individual subjects when measured using long-range correlations [2]. The output of a control system reflects the underlying control system’s fluctuations as they act on feedback loops [37].
The model output generated a cyclic pattern with breath-to-breath variations that mimicked VT’s properties when noise was added to the neural network model of the brain respiratory oscillator [38]. Cycle-by-cycle variability impacts the functioning of cells in the respiratory system [2]. Sleep, pulmonary diseases, hypoxia, and anxiety disorders affect breathing variability [2]. Variability also occurs in resting states, suggesting that it is inherent to normal pulmonary function.
The variability and correlation properties of spontaneous breathing in humans were investigated in seated healthy subjects. α represents the scaling exponent of a power-law fluctuation function. For a random process, α has a value of 0.5. A positively correlated signal shows large fluctuations following large fluctuations. A signal with small fluctuations following large fluctuations has a value between 0 and 0.5 and is considered anti-correlated [2]. Breath-by-breath, the VT, respiratory time (TTOT), and other breathing pattern parameters were computed. Regarding minute ventilation and respiratory rate, VT, TTOT, and other breathing variables showed values between 0.60 and 0.71 [2].
The brain’s respiratory center is one of the regulatory systems for fluctuations in breathing parameters; it receives input from other brain centers. Except for the end-expiratory lung volume (EEVL), which is related to tissue viscoelasticity, all respiratory variables are correlated with the brain [2]. Correlated fluctuations originate in the brain, but periphery input is necessary to maintain the oscillations [39]. The oscillator’s output would be a deterministic cyclic pattern with constant frequency and amplitude without fluctuations in the inputs to the neural oscillator [40]. While the correlated properties of EEVL are influenced by the brain oscillator and viscoelastic memory of the tissues, breathing patterns are influenced by the brain and neural noise [2,41].
Oscillators are non-linear networks of neuron groups, so lung and other brain area fluctuations can produce cycle-by-cycle variations in the output, which may be influenced by respiratory muscles [41]. Stretch-sensitive adherent cells in the respiratory system are affected by cycle-by-cycle fluctuations in tidal volume and respiratory time [42]. As a result of this variability, respiratory system cells adapt [2].
A reduction in the active brain control of respiratory variables occurs during sleep [43]. Brainstem catecholamine (CA) neurons manifest an arousal-state-dependent activity pattern and participate in the control of breathing while modulating the processing of sensory information [44,45]. CA neurons are involved in central respiratory chemoreception, which boosts breathing frequency and lowers breathing variations during rapid eye movement (REM) sleep [46].
Per the CDP, respiratory variability is inherent to normal cellular respiration and lung function and characterizes brain centers, lungs, and respiratory cells. Although it can adapt to the changing environment, variability also characterizes the resting state, implying that the respiratory system requires it for proper function [4].
The key to understanding how the inherent variability in breathing and respiration contributes to the system’s function and efficiency is to be determined. Another unresolved issue is to what extent this variability is personalized and to what extent learning and systems’ history impact the variability.

1.3. The Regulation of Cellular Respiration and Electrons Transport

As a result of cellular respiration, free energy can be released from carbohydrate, fat, and protein energy substrates in a controlled manner [47,48]. While the sequence seems to follow a particular ordered path, variability underlies many of these processes.
In the cytosol, glycolysis occurs, while in the mitochondria, the citric acid cycle occurs, along with oxidative phosphorylation [49]. It requires glucose, adenosine triphosphate (ATP), and NAD+; its products are ATP and H2O [50]. Oxidative phosphorylation involves the electron transport chain and chemiosmosis [51]. Electrons travel along the transport chain’s components through redox reactions. The energy of these reactions release is captured as a proton gradient, which makes ATP and is termed chemiosmosis [24,52].
In addition to proton leakage, electron leakage from electron transport chain complexes uncouples membrane voltage from ATP production. If electrons leak prematurely, they pass directly to oxygen rather than through oxygen to form water, causing superoxide to form, which can be toxic to cells [53]. This process suggests a more complicated regulatory mechanism involving a degree of variability.
There are several regulatory steps along the respiration pathway. An enzyme that catalyzes the pathway’s first committed step is the main target for regulating biochemical processes [53]. When an enzyme is active, a specific step proceeds; if it is inactive, the step may not occur or will be slowed. The binding of regulatory molecules ATP, ADP, and NADH controls respiration enzymes at the allosteric sites. Increased ATP level implies that a cell has sufficient ATP, leading to inhibition of the respiratory system [54].
A vital control point is a phosphofructokinase (PFK). Adenosine monophosphate (AMP) is a positive regulator of PFK, while ATP is a negative regulator [55]. By converting pyruvate to acetyl CoA, pyruvate dehydrogenase controls the acetyl CoA entering citric acid. When ATP and NADH are present, this enzyme is less active; when ADP is present, it is more active. Pyruvate dehydrogenase is also activated by its substrate, pyruvate, and inhibited by its product, acetyl CoA [56].
Pyruvate dehydrogenase regulates citric acid cycle entry. Additional regulatory points include the release of carbon dioxide molecules and the production of the first two NADH molecules of the cycle. ADP and ATP inhibit isocitrate dehydrogenase, while NADH activates it. ATP, NADH, and several other molecules, including succinyl CoA, inhibit α-ketoglutarate dehydrogenase [57,58].
Additional regulatory mechanisms are the oxygen delivery to tissues, regulation of oxygen binding to the heme moiety of cytochrome oxidase by NO, availability of nutrient metabolism to generate NADH and FADH2, and the overall cellular energy state defined by the ratio of ATP/ADP [57,58,59].
By oxidative phosphorylation, mitochondria coordinate energy demands with energy production in the cytosol. Calcium signals between the cytosol and matrix regulate this process. Transfer of Ca2+ into the mitochondria signals an increased energy demand. Upregulating citric acid cycle enzymes, increasing respiration, and increasing ATP synthesis increases energy provision by oxidative phosphorylation [60,61]. Changes in mitochondrial morphology impact the bioenergetic state, and changes in bioenergetics result in altered morphology [62].
These control mechanisms underlie the ordered process of cellular respiration but do not explain the inherent variability of the processes.

1.4. The Constrained Disorder Principle Accounts for Variability in Cellular Respiration

Variation in respiratory functions serves as a mechanism for systems’ flexibility and adaptation to environmental changes [63]. The respiratory system acclimates over time to the environment by adjusting its enzyme capacities. Adaptation is the genetic divergence of the properties of the respiratory system [63]. Environmental factors induce variation in mitochondrial efficiency that impacts performance.
Living organisms adapt to local environmental conditions through natural selection of genetic variation. Genetic variation in mitochondrial efficiency and oxygen consumption is relevant for energy metabolism. Over time, mitochondria produce ATP at different rates in different subjects, populations, and environments within the same individual [64,65,66,67]. Throughout an individual’s lifetime, the degree of energy coupling derived from oxidation varies across tissues, among individuals, and over time [68,69].
Measuring ATP production from oxygen consumption is inaccurate because the amount produced per unit of oxygen consumed varies significantly [70]. Because of the inherent variability in the link between oxidation and phosphorylation and the amount of ATP generated per molecule of oxygen consumed by mitochondria (P/O ratio), oxygen consumption is not a reliable marker of energy metabolism. Diet and temperature affect the P/O ratio within and among subjects. Due to tissue-specific regulation of mitochondria, mitochondrial function and P/O ratio differ significantly among tissues within the same subject [71,72]. The variability of the P/O ratio impacts an organism’s performance. Reducing the P/O ratio is energetically costly but provides advantages via reduced production of reactive oxygen species [73,74].
The relationship between the P/O ratio and performance is not causative [75]. An optimal P/O ratio and oxygen consumption rate are shaped by extrinsic (food availability and temperature) and intrinsic variables (genotype, hormones, and substrate mobilization) [73]. A population, individual, and adaptive strategy for maximizing mitochondrial energy efficiency and operating at the lowest oxidative cost is necessary [73]. It is possible to expect an increase in the P/O ratio during periods of high energy demand, even if this results in an increase in reactive oxygen species (ROS) [76].
Variation in the P/O ratio involves proton leakage across the membrane, contributing to the drop in the electrochemical potential (Δp) and oxygen consumption, independent of ATP production [77,78]. Several factors affect mitochondrial membrane conductance, including phospholipid fatty acids and mitochondrial carrier proteins, such as the uncoupling protein (UCP) and the adenine nucleotide transporter [79]. The Δp is affected by the active transport across the inner membrane of anions (ADP3− and ATP4−), cations (Ca2+), and metabolites (aspartate and glutamate) [73,80].
Temperature affects ATP production due to an increase in the H+ conductance of the inner membrane. At higher temperatures, more H+ is shunted away from ATP synthase, reducing P/O [81]. Low temperatures reduce mitochondrial coupling in brown adipose tissue, contributing to thermogenesis [82]. Acclimation to cold temperatures increases oxygen consumption but does not change the coupling of ATP production to oxygen consumption [83].
Stress compensates for lowered respiratory ATP production efficiency with higher respiratory rates [84]. Respiratory supply and demand changes accompany stress and mask the effects of lowered efficiency [85]. Oxidative phosphorylation improves respiration in low-oxygen environments. A study of fish populations with different temperature regimes and their gene expression of mitochondrial-encoded oxidative phosphorylation subunits identified seven transcripts showing increased expression levels that differ with changing temperatures, improving respiratory efficiency [86].
By reducing food intake, the P/O ratio increases, reducing ATP synthesis costs and energy substrate requirements [66,87]. As a result of diet, phospholipid properties of the inner mitochondrial membrane are altered, affecting the P/O ratio. Unsaturated fats increase protons’ permeability and mitochondrial proton leak, altering the P/O ratio [88].
Evolution was expected to maximize the P/O ratio for improved resource usage. However, natural selection did not maximize mitochondrial efficiency due to the generation of ROS. The variation in the P/O ratio may be because of the ATP production trade-off against ROS generation that results from a flow of electrons through the mitochondrial ETC [73,89,90,91,92]. Reducing the P/O ratio may slow aging via a reduction in ROS production [93]. By lowering mitochondrial respiration, partial pressure of oxygen build-up increases ROS production and ATP synthesis and accelerates respiration and electron flow through the ETC [94].
A cell’s ability to generate ATP determines its growth [95]. Growth efficiency varies between subjects depending on mitochondrial function. In individuals with higher growth efficiencies, UCP expression is reduced, proton leak is reduced, and P/O ratios are higher [96,97].
In plants, respiratory variability results from genetic, developmental, and environmental factors [63]. It optimizes respiration’s energetic and biosynthetic performance under changing conditions [63]. Physiological supply and demand conditions determine the rate at which substrates enter and leave the respiratory system. The flexibility of the process is reflected in variations in how carbon substrates are metabolized. Although substrate levels do not directly affect respiratory enzymes, they can indirectly stimulate respiration. A tissue’s respiration rate is regulated by respiratory enzymes whose regulatory properties are targeted by signaling based on energy status [98,99,100]. By engaging alternate routes and increasing the permeability of the mitochondrial inner membrane to protons, every ATP-producing reaction in respiration can be bypassed, resulting in a variation in ATP yield [101].
The inherent variability of the respiratory system is a challenge for modeling the regulatory control of the system.

1.5. Tunneling in Redox Reactions Implies That Variability Underlies Respiration at the Atomic Level and Is a Manifestation of Quantum Effects

While the sequences of cellular respiration and some of its variability are well determined, the classical approach cannot explain the atomic events and their regulation. From a thermodynamic perspective, a cell is a semi-open system that allows energy to enter and waste to leave [102,103]. If a cell’s energy levels fall, it disrupts the balance of order and disorder. It is because it has too much or lacks sufficient disorder, affecting the cell’s function and leading to cellular death [103]. According to the CDP, the cell requires energy to maintain order and disorder [24]. It implies that energy is required to maintain dynamic boundaries for systems’ function.
There is randomness at the core of the quantum world of atoms and particles [104]. Calculating probabilistic quantum wave functions entails randomness playing a role in this effect [105]. Tunneling, coherence, and entanglement are linked to DNA, which is the blueprint for respiratory chain proteins; it follows that quantum physics is linked with diseases of the respiratory chain [3].
During a redox reaction, electrons are transferred between two chemical entities simultaneously: oxidation (loss of electrons) and reduction (gain of electrons) [106]. Molecular pumps, concentration gradients across membranes, and energy-rich metabolites are all regulated by redox reactions. Alterations in redox balance contribute to aging and disease progression as the genome adapts to environmental challenges [107,108,109].
Proton transfer is crucial in several enzyme-catalyzed reactions, which involve the movement of protons between molecules and may involve quantum-mechanical tunneling, whereby a particle passes through an energy barrier rather than gathering energy to ‘climb’ over it [110]. Proton (hydrogen) tunneling plays a role in numerous enzyme reactions [110,111]. The proton-coupled electron transfer process involves long-range electron tunneling, and hydrogen tunneling is widespread in biological processes, including DNA repair, photosynthesis, cellular homeostasis, and cell death, being required for complex chemical transformations [112,113].
By connecting ROS partitioning and cellular bioenergetics, quantum effects bridge the atomic and cellular levels [114]. Coherent electron spin dynamics influence ROS production. Reduced flavoenzymes formed spin-correlated radical pairs during molecular oxygen activation in cell culture. Cellular superoxide and hydrogen peroxide ROS products are altered by oscillating magnetic fields, indicating coherent singlet-triplet mixing. The orientation dependence of magnetic stimulation involves changes in ROS levels, increasing mitochondrial respiration and glycolysis rates [114]. The radical pair mechanism determines how quantum effects affect ROS production. Electron-nuclear hyperfine interactions, internal magnetic interactions, and applied magnetic fields control spin dynamics [115].
An electron transfer activates oxygen in reduced flavoenzymes, resulting in magnetically sensitive ROS formation [116]. Spin-correlated radical pairs are formed between flavin semiquinone (FADH) and superoxide (O2•−) [117]. Coherent evolution between the singlet and triplet states of FADH:O2•− radical pairs determines the products of the reaction [118]. ROS production is differentially affected at parallel excitation, indicating increased mitochondrial respiration [119,120,121,122].
The data support the notion that quantum mechanics is essential and that randomness plays a role in the proper respiration process [123]. The CDP applies to quantum effects and thus defines both classical and quantum effects in respiration [4].

1.6. Altered Variability in Lung Diseases

The CDP implies that chronic lung diseases are associated with deviations in breathing pattern variability [124,125,126]. Altered variability in physiological and cellular characteristics in different pulmonary diseases can predict and possibly prevent clinical deterioration [127]. Additionally, variability can be a valuable tool for diagnosis and treatment adjustments in those patients.
Respiratory variability is lower in chronic obstructive pulmonary disease (COPD) than in healthy patients. It may be due to lung and chest wall mechanics changes or neural adjustments in breathing control [31]. COPD reduces the variation in inter-breathing intervals that characterize normal breathing and is influenced by metabolic demands [128]. Patients with COPD have more regular breath-to-breath fluctuations than those without COPD, and the increased regularity is correlated with the severity of the disease. Compared to healthy controls, patients with COPD need to breathe faster to compensate, which increases oxygen demand [128].
Patients with restrictive lung disease display breathing variability significantly smaller than healthy patients [129]. The variability is significant for tidal volume and expiratory time. It increases dyspnea in response to slight variations from the average resting tidal volume. During lung disease of infancy, non-REM sleep, and highly demanding cognitive tasks, breathing variability is reduced [31]. Variability in tidal breathing parameters is a risk factor for developing subsequent respiratory morbidity in preterm infants [129]. During the first year of life, preterm infants with low variability in tidal breathing parameters were more likely to be re-hospitalized [129].
In contrast to patients with COPD and interstitial lung disease (ILD), asthmatics’ respiration variability increases with the disease’s severity [31]. The quantitative variability of tidal breathing parameters in children with asthma was increased and improved but not normalized following bronchodilators [130]. Hypoxia, hypertension, and anxiety disorders increase the variability of older adults during complex arithmetic tasks [31]. It is also affected by the use of different anesthetics and sedation drugs.
Non-invasive pulmonary function tests (PFTs) identify daily and day-to-day variability. Peak expiratory flow (PEF) variability measures are reliable predictors of loss of asthma control and response to treatment [131,132,133]. The quantitative variability of minute ventilation and tidal volume is higher in COPD patients than in age- and sex-matched controls, and sighs are significantly reduced in COPD patients [134]. Another method to evaluate variability in airway resistance is using the forced oscillation technique (FOT) [135]. This method predicts asthma exacerbations [135,136]. In COPD, there is similar high day-to-day variability in FOT [137]. FOT variability is related to symptoms and precedes them by a few days and can detect changes before COPD exacerbation, enabling preemptive measures [138]. Patients with ILD show significant variability in forced vital capacity (FVC) during the day. This variability increases in patients with progressive disease compared to stable ones. FVC variability may be used in patients with ILD as a disease progression predictor [139].
Subjects receiving an endotoxin challenge, as well as postoperative and acutely ill patients, have reduced breathing pattern variability [140,141]. All organs, including the respiratory center, arterial chemoreceptors, lung vagal sensory receptors, and lung mechanics regulating breathing variability, are affected by endotoxin [142,143].
These data follow the CDP, implying that reduced or out-of-bounds high degrees of variability are associated with reduced systems’ efficiency.

1.7. Using the Constrained Disorder Principle-Based Platform for Augmenting Cellular Respiration and Improving Therapies for Chronic Respiratory Diseases

Variability characterizes the proper function of the respiratory system. Per the CDP, loss of variability or increased variability outside borders is associated with a diseased system [4]. Developing a CDP-based second-generation AI system, which implements algorithms based on variability, can correct system malfunctions and augment efficiency, increasing the variability in cases where it is too low or controlling it in cases where it is out of control. Based on the CDP, these measures are expected to improve clinical conditions by overcoming drug tolerance in patients with chronic lung conditions [4,144].
The algorithm can be implemented for the respiratory system to improve medication response in patients with chronic lung disorders who develop tolerance to therapies. By measuring the reaction to beta-agonists following bronchoconstriction with methacholine, it has been shown that regular beta-agonists use leads to tolerate their bronchodilator effects [145]. The bronchodilator response to salbutamol is reduced in patients taking formoterol due to tolerance. Despite increased bronchoconstriction and reduced FEV1, regular salbutamol attenuated the acute response to agonists. Increasing bronchoconstriction conferred increased susceptibility to the effects of bronchodilator tolerance [146,147].
The second-generation AI system is based on three levels. At the first level, the system implements variability in the output therapeutic regimen independent of the input. It provides a randomized medication regimen for an inhaler regarding dosage and administration time within a pre-defined approved range [21,22,23]. At the second level, a closed-loop system is implemented, wherein the system collects data on the output. The clinically meaningful endpoints are used as continuous inputs for the algorithm, altering the randomization to reach a better outcome. The algorithm personalizes the output based on the inputs it receives [4,19,21,22,23,144,148,149,150,151,152,153,154,155,156,157,158,159,160,161,162,163,164,165,166,167,168,169,170,171,172]. At the third level, the algorithm receives inputs from the user’s quantifications of variability signatures. These include measurements of the variability of different breathing parameters or variability in cytokines measurements. The algorithm uses large datasets from big data resources on patients with similar lung conditions while updating them and personalizing them to the individual patient [19,148,149,150,151,152,153,154,155,156,157,158,159,160,161,162,163,164,165].
Figure 1 shows a schematic presentation of the three steps for overcoming tolerance to bronchodilators by implementing variability-based therapeutic regimens.

1.8. The CDP-Based Platform for Improving Mechanical Ventilation and Assessment of Extubation Readiness

The CDP-based second-generation AI system can improve mechanical ventilation’s effectiveness, ease weaning from devices, and overcome drug tolerance in patients with chronic lung diseases by implementing variabilities.
Fixed breathing frequencies and monotonous tidal volumes characterize current ventilation methods. Using this non-physiological method of breathing ignores breathing variability and may damage the lungs, causing them to collapse and lowering the partial pressure of oxygen in the arteries to abnormally low levels [173]. To address this issue, ‘noisy ventilation’ was developed using biologically variable computer-controlled ventilators that incorporate noise to mimic natural variability. Opening collapsed alveoli without damaging them may be possible by occasional high pressures [174]. ‘Noisy ventilation’ implements breathing variability and improves oxygenation in animals with low mean airway pressure, comparable to minute ventilation [175].
Variable ventilation was tested in several preclinical studies. In a porcine model of oleic-acid-induced acute lung injury, animals received variable ventilation, which included changing breathing frequency in the range of 15–27 breaths/min and variable VT or conventional ventilation at a rate of 20 breaths/min. There was a significant improvement in arterial oxygenation and lower shunt fraction in the variable ventilation group [176]. Compared to conventional mechanical ventilation, variable ventilation led to a twofold increase in bronchoalveolar lavage phospholipid surfactant levels and reduced alveolar protein content in guinea pigs [177]. If variable ventilation is combined with PEEP, recruitment is more sustained than monotonous controlled mechanical ventilation [178,179]. VT fluctuation at a frequency of 0.05 Hz improved oxygenation during variable ventilation in an experimental model of severe acute respiratory distress syndrome (ARDS) [42]. Variable ventilation may be further enhanced by deterministic resonance. Mechanical ventilation improves gas exchange and lung mechanics with varying VT in ARDS [180,181]. As a result, ventilator-induced lung injury (VILI) is reduced [182]. Variable ventilation facilitates uniform recruitment of the lungs without enlarging hyperaerated lung areas during acute lung injury [178,183].
During elective abdominal aortic aneurysmectomy, 41 patients were randomized to conventional or variable ventilation. With variable ventilation, arterial oxygenation and pulmonary compliance were higher with lower PaCO2 and dead-space ventilation [184]. The variable ventilation model also succeeded in patients with other lung injuries besides ARDS [185].
Both unnecessarily long extubation times and premature extubation are linked with adverse outcomes [186]. Assessment of extubation readiness for invasively ventilated patients is associated with a substantial failure rate [187,188]. Variability in breathing also serves as a predictor of successful extubation [189]. Altered respiratory rate variability and reduced complexity are reliable predictive tools for failed extubation in adults [190]. When applied to basic respiratory parameters, breathing variability predicts extubation failure, and comprehensive breathing variability indices improve the prediction. Based on continuous ventilation waveforms, breath-to-breath basic and comprehensive respiratory parameters were computed an hour before extubation [188]. During spontaneous breathing, intermittent deep-inflation or sigh breaths are characterized by large VT breaths (2–3 times standard VT) [191]. The sigh breath airs collapsed alveoli, improves full residual capacity (FRC), reduces pulmonary shunt, and resets the breathing pattern in excessive or unvarying breathing patterns [191]. Applying sigh breaths to ventilated hypoxemic patients improves ventilator-free days [192,193].
Weaning failure during mechanical ventilation is predicted by low respiratory variability. Patients who failed weaning trials had higher variability in spontaneous tidal volume and peak inspiratory flow [194]. In patients who failed to wean after cardiac surgery, the pattern of tidal volume, but not the respiratory rate, was more irregular due to weaning, and hemodynamic and autonomic nervous system changes occur [32]. In postoperative patients recovering from systemic inflammatory response syndrome (SIRS), breathing pattern variability was a weaning predictor [142]. For 30 min before weaning, tidal volume, total breath duration, inspiratory time, expiratory time, and peak inspiratory flow were monitored in mechanically ventilated SIRS patients. The five parameters’ average values of total breath duration, inspiratory time, expiratory time, and coefficient of variation were lower in the failure group, suggesting that SIRS failure groups were less likely to maintain breathing pattern variability due to reduced regulatory mechanisms [142]. After 60 min of spontaneous breathing before extubation, ventilated patients successfully separated from the ventilator showed greater breathing variability [189]. A CV of tidal volume/inspiratory time ≥ 19% and a CV of inspiratory time/respiratory period ≥ 10% were used to determine whether extubation was successful or unsuccessful [189]. A reduction in heart rate variability during spontaneous breathing trials (SBT) was associated with extubation failure, demonstrating the importance of variability in different organs to maintain proper function [190,195]. In contrast, high variability in breathing patterns in children was associated with failed extubation [196].
Figure 2 shows a schematic presentation of inserting variability into mechanical ventilation to improve the efficacy of ventilation and ease the weaning process.

2. Summary

Variability characterizes breathing and cellular respiration. Using variability by implementing disorder-based algorithms can improve the response to drugs in patients with chronic lung disorders and improve ventilation techniques. Personalization of variability-based algorithms and the linkage between the degree of variability and outcome are some of the future challenges faced by these platforms. Ongoing clinical studies are expected to shed light on the application of these measures in different clinical settings.

Author Contributions

O.A., A.H. and Y.I. wrote the manuscript. Y.I. conceptualized the study. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Data Availability Statement

All data are available on public domains.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

Constrained disorder principle: CDP; artificial intelligence: AI; tidal volume: VT; respiratory time: TTOT; end-expiratory lung volume: EEVL; catecholamine: CA; rapid eye movement: REM; adenosine triphosphate: ATP; phosphofructokinase: PFK; adenosine monophosphate: AMP; reactive oxygen species: ROS; electrochemical potential: Δp; uncoupling protein: UCP; chronic obstructive pulmonary disease: COPD; interstitial lung disease: ILD; pulmonary function tests: PFTs; peak expiratory flow: PEF; forced oscillation technique: FOT; forced vital capacity: FVC; acute respiratory distress syndrome: ARDS; ventilator-induced lung injury: VILI; full residual capacity: FRC; systemic inflammatory response syndrome: SIRS; closing volume: CV; spontaneous breathing trials: SBT.

References

  1. Patel, H.; Kerndt, C.C.; Bhardwaj, A. Physiology, Respiratory Quotient; StatPearls Publishing: Treasure Island, FL, USA, 2022. [Google Scholar]
  2. Dellaca, R.L.; Aliverti, A.; Mauro, A.L.; Lutchen, K.R.; Pedotti, A.; Suki, B. Correlated Variability in the Breathing Pattern and End-Expiratory Lung Volumes in Conscious Humans. PLoS ONE 2015, 10, e0116317. [Google Scholar] [CrossRef]
  3. Calvillo, L.; Redaelli, V.; Ludwig, N.; Qaswal, A.B.; Ghidoni, A.; Faini, A.; Rosa, D.; Lombardi, C.; Pengo, M.; Bossolasco, P.; et al. Quantum Biology Research Meets Pathophysiology and Therapeutic Mechanisms: A Biomedical Perspective. Quantum Rep. 2022, 4, 148–172. [Google Scholar] [CrossRef]
  4. Ilan, Y. The constrained disorder principle defines living organisms and provides a method for correcting disturbed biological systems. Comput. Struct. Biotechnol. J. 2022, 20, 6087–6096. [Google Scholar] [CrossRef]
  5. Finn, E.H.; Misteli, T. Molecular basis and biological function of variability in spatial genome organization. Science 2019, 365, eaaw9498. [Google Scholar] [CrossRef]
  6. Chiera, M.; Cerritelli, F.; Casini, A.; Barsotti, N.; Boschiero, D.; Cavigioli, F.; Corti, C.G.; Manzotti, A. Heart Rate Variability in the Perinatal Period: A Critical and Conceptual Review. Front. Neurosci. 2020, 14, 561186. [Google Scholar] [CrossRef]
  7. Forte, G.; Favieri, F.; Casagrande, M. Heart Rate Variability and Cognitive Function: A Systematic Review. Front. Neurosci. 2019, 13, 710. [Google Scholar] [CrossRef]
  8. Mitchison, T.; Kirschner, M. Dynamic instability of microtubule growth. Nature 1984, 312, 237–242. [Google Scholar] [CrossRef]
  9. Kirschner, M.W.; Mitchison, T. Microtubule dynamics. Nature 1986, 324, 621. [Google Scholar] [CrossRef] [PubMed]
  10. Ilan, Y. Overcoming randomness does not rule out the importance of inherent randomness for functionality. J. Biosci. 2019, 44, 1–12. [Google Scholar] [CrossRef]
  11. Ilan, Y. Generating randomness: Making the most out of disordering a false order into a real one. J. Transl. Med. 2019, 17, 1–12. [Google Scholar] [CrossRef] [PubMed]
  12. Ilan, Y. Advanced Tailored Randomness: A Novel Approach for Improving the Efficacy of Biological Systems. J. Comput. Biol. 2020, 27, 20–29. [Google Scholar] [CrossRef]
  13. Ilan, Y. Order Through Disorder: The Characteristic Variability of Systems. Front. Cell Dev. Biol. 2020, 8, 186. [Google Scholar] [CrossRef]
  14. El-Haj, M.; Kanovitch, D.; Ilan, Y. Personalized inherent randomness of the immune system is manifested by an individualized response to immune triggers and immunomodulatory therapies: A novel platform for designing personalized immunotherapies. Immunol. Res. 2019, 67, 337–347. [Google Scholar] [CrossRef] [PubMed]
  15. Ilan, Y. Randomness in microtubule dynamics: An error that requires correction or an inherent plasticity required for normal cellular function? Cell Biol. Int. 2019, 43, 739–748. [Google Scholar] [CrossRef] [PubMed]
  16. Ilan, Y. Microtubules: From understanding their dynamics to using them as potential therapeutic targets. J. Cell. Physiol. 2018, 234, 7923–7937. [Google Scholar] [CrossRef] [PubMed]
  17. Ilan-Ber, T.; Ilan, Y. The role of microtubules in the immune system and as potential targets for gut-based immunotherapy. Mol. Immunol. 2019, 111, 73–82. [Google Scholar] [CrossRef]
  18. Forkosh, E.; Kenig, A.; Ilan, Y. Introducing variability in targeting the microtubules: Review of current mechanisms and future directions in colchicine therapy. Pharmacol. Res. Perspect. 2020, 8, e00616. [Google Scholar] [CrossRef] [PubMed]
  19. Ilan, Y. β-Glycosphingolipids as Mediators of Both Inflammation and Immune Tolerance: A Manifestation of Randomness in Biological Systems. Front. Immunol. 2019, 10, 1143. [Google Scholar] [CrossRef] [PubMed]
  20. Ilan, Y. Microtubules as a potential platform for energy transfer in biological systems: A target for implementing individualized, dynamic variability patterns to improve organ function. Mol. Cell. Biochem. 2022, 478, 375–392. [Google Scholar] [CrossRef] [PubMed]
  21. Ilan, Y. Overcoming Compensatory Mechanisms toward Chronic Drug Administration to Ensure Long-Term, Sustainable Beneficial Effects. Mol. Ther.-Methods Clin. Dev. 2020, 18, 335–344. [Google Scholar] [CrossRef]
  22. Ilan, Y. Second-Generation Digital Health Platforms: Placing the Patient at the Center and Focusing on Clinical Outcomes. Front. Digit. Heal. 2020, 2, 569178. [Google Scholar] [CrossRef]
  23. Ilan, Y. Improving Global Healthcare and Reducing Costs Using Second-Generation Artificial Intelligence-Based Digital Pills: A Market Disruptor. Int. J. Environ. Res. Public Heal. 2021, 18, 811. [Google Scholar] [CrossRef]
  24. Ilan, Y. Enhancing the plasticity, proper function and efficient use of energy of the Sun, genes and microtubules using variability. Clin. Transl. Discov. 2022, 2, e103. [Google Scholar] [CrossRef]
  25. Tobin, M.J.; Mador, M.J.; Guenther, S.M.; Lodato, R.F.; Sackner, M.A.; Gustafsson, P.M.; Bengtsson, L.; Lindblad, A.; Robinson, P.D.; Suki, B.; et al. Variability of resting respiratory drive and timing in healthy subjects. J. Appl. Physiol. 1988, 65, 309–317. [Google Scholar] [CrossRef] [PubMed]
  26. Benchetrit, G. Breathing pattern in humans: Diversity and individuality. Respir. Physiol. 2000, 122, 123–129. [Google Scholar] [CrossRef] [PubMed]
  27. Napoli, N.J.; Rodrigues, V.R.; Davenport, P.W. Characterizing and Modeling Breathing Dynamics: Flow Rate, Rhythm, Period, and Frequency. Front. Physiol. 2022, 12, 772295. [Google Scholar] [CrossRef] [PubMed]
  28. Brack, T.; Jubran, A.; Tobin, M.J. Effect of Resistive Loading on Variational Activity of Breathing. Am. J. Respir. Crit. Care Med. 1998, 157, 1756–1763. [Google Scholar] [CrossRef] [PubMed]
  29. Brack, T.; Jubran, A.; Tobin, M.J. Effect of elastic loading on variational activity of breathing. Am. J. Respir. Crit. Care Med. 1997, 155, 1341–1348. [Google Scholar] [CrossRef] [PubMed]
  30. Jubran, A.; Grant, B.J.B.; Tobin, M.J. Effect of Hyperoxic Hypercapnia on Variational Activity of Breathing. Am. J. Respir. Crit. Care Med. 1997, 156, 1129–1139. [Google Scholar] [CrossRef] [PubMed]
  31. Bosch, O.F.C.v.D.; Alvarez-Jimenez, R.; de Grooth, H.-J.; Girbes, A.R.J.; Loer, S.A. Breathing variability—Implications for anaesthesiology and intensive care. Crit. Care 2021, 25, 1–13. [Google Scholar] [CrossRef]
  32. Engoren, M. Approximate entropy of respiratory rate and tidal volume during weaning from mechanical ventilation. Crit. Care Med. 1998, 26, 1817–1823. [Google Scholar] [CrossRef] [PubMed]
  33. Fishman, M.; Jacono, F.J.; Park, S.; Jamasebi, R.; Thungtong, A.; Loparo, K.A.; Dick, T.E. A method for analyzing temporal patterns of variability of a time series from Poincaré plots. J. Appl. Physiol. 2012, 113, 297–306. [Google Scholar] [CrossRef] [PubMed]
  34. Kamen, P.W.; Tonkin, A.M. Application of the Poincaré plot to heart rate variability: A new measure of functional status in heart failure. Aust. N. Z. J. Med. 1995, 25, 18–26. [Google Scholar] [CrossRef]
  35. Fadel, P.J.; Barman, S.M.; Phillips, S.W.; Gebber, G.L. Fractal fluctuations in human respiration. J. Appl. Physiol. 2004, 97, 2056–2064. [Google Scholar] [CrossRef] [PubMed]
  36. Campana, L.; Owens, R.; Butler, J.; Suki, B.; Malhotra, A. Variability of respiratory mechanics during sleep in overweight and obese subjects with and without asthma. Respir. Physiol. Neurobiol. 2013, 186, 290–5. [Google Scholar] [CrossRef]
  37. Glass, L. Synchronization and rhythmic processes in physiology. Nature 2001, 410, 277–284. [Google Scholar] [CrossRef] [PubMed]
  38. Cernelc, M.; Suki, B.; Reinmann, B.; Hall, G.L.; Frey, U. Correlation properties of tidal volume and end-tidal O2 and CO2 concentrations in healthy infants. J. Appl. Physiol. 2002, 92, 1817–1827. [Google Scholar] [CrossRef]
  39. Botros, S.M.; Brace, E.N. Neural network implementation of a three-phase model of respiratory rhythm generation. Biol. Cybern. 1990, 63, 143–153. [Google Scholar] [CrossRef] [PubMed]
  40. Wilson, C.J. Oscillators and Oscillations in the Basal Ganglia. Neurosci. 2014, 21, 530–539. [Google Scholar] [CrossRef] [PubMed]
  41. Betka, S.; Adler, D.; Similowski, T.; Blanke, O. Breathing control, brain, and bodily self-consciousness: Toward immersive digiceuticals to alleviate respiratory suffering. Biol. Psychol. 2022, 171, 108329. [Google Scholar] [CrossRef]
  42. Güldner, A.; Huhle, R.; Beda, A.; Kiss, T.; Bluth, T.; Rentzsch, I.; Kerber, S.; Carvalho, N.C.; Kasper, M.; Pelosi, P.; et al. Periodic Fluctuation of Tidal Volumes Further Improves Variable Ventilation in Experimental Acute Respiratory Distress Syndrome. Front. Physiol. 2018, 9, 905. [Google Scholar] [CrossRef] [PubMed]
  43. Chowdhuri, S.; Badr, M.S. Control of Ventilation in Health and Disease. Chest 2016, 151, 917–929. [Google Scholar] [CrossRef] [PubMed]
  44. Li, A.; Emond, L.; Nattie, E. Brainstem Catecholaminergic Neurons Modulate both Respiratory and Cardiovascular Function. In Integration in Respiratory Control; Springer: New York, NY, USA, 2008; Volume 605, pp. 371–376. [Google Scholar] [CrossRef]
  45. Guyenet, P.G.; Bayliss, D.A. Neural Control of Breathing and CO2 Homeostasis. Neuron 2015, 87, 946–961. [Google Scholar] [CrossRef]
  46. Li, A.; Nattie, E. Catecholamine neurones in rats modulate sleep, breathing, central chemoreception and breathing variability. J. Physiol. 2006, 570, 385–396. [Google Scholar] [CrossRef]
  47. Osellame, L.D.; Blacker, T.S.; Duchen, M.R. Cellular and molecular mechanisms of mitochondrial function. Best Pract. Res. Clin. Endocrinol. Metab. 2012, 26, 711–723. [Google Scholar] [CrossRef]
  48. Zimmerman, J.J.; von Saint André-von Arnim, A.; McLaughlin, J. Cellular Respiration. In Pediatric Critical Care, 4th ed.; Fuhrman, B.P., Zimmerman, J.J., Eds.; Mosby: Saint Louis, MO, USA, 2011; Chapter 74; pp. 1058–1072. [Google Scholar]
  49. Martínez-Reyes, I.; Chandel, N.S. Mitochondrial TCA cycle metabolites control physiology and disease. Nat. Commun. 2020, 11, 102. [Google Scholar] [CrossRef] [PubMed]
  50. Yang, Y.; Sauve, A.A. NAD + metabolism: Bioenergetics, signaling and manipulation for therapy. Biochim. et Biophys. Acta (BBA)-Proteins Proteom. 2016, 1864, 1787–1800. [Google Scholar] [CrossRef] [PubMed]
  51. Manoj, K.M. Aerobic Respiration: Criticism of the Proton-centric Explanation Involving Rotary Adenosine Triphosphate Synthesis, Chemiosmosis Principle, Proton Pumps and Electron Transport Chain. Biochem. Insights 2018, 11. [Google Scholar] [CrossRef]
  52. Milshteyn, D.; Cooper, G.; Deamer, D. Chemiosmotic energy for primitive cellular life: Proton gradients are generated across lipid membranes by redox reactions coupled to meteoritic quinones. Sci. Rep. 2019, 9, 1–12. [Google Scholar] [CrossRef]
  53. Robinson, P.K. Enzymes: Principles and biotechnological applications. Essays Biochem. 2015, 59, 1–41. [Google Scholar] [CrossRef]
  54. Li, M.; Li, C.; Allen, A.; Stanley, C.A.; Smith, T.J. The structure and allosteric regulation of glutamate dehydrogenase. Neurochem. Int. 2011, 59, 445–455. [Google Scholar] [CrossRef]
  55. McGresham, M.S.; Lovingshimer, M.; Reinhart, G.D. Allosteric Regulation in Phosphofructokinase from the Extreme Thermophile Thermus thermophilus. Biochemistry 2013, 53, 270–278. [Google Scholar] [CrossRef] [PubMed]
  56. Sugden, M.C.; Holness, M.J. The pyruvate carboxylase-pyruvate dehydrogenase axis in islet pyruvate metabolism: Going round in circles? Islets 2011, 3, 302–319. [Google Scholar] [CrossRef]
  57. Wilson, D.F. Oxidative phosphorylation: Regulation and role in cellular and tissue metabolism. J. Physiol. 2017, 595, 7023–7038. [Google Scholar] [CrossRef] [PubMed]
  58. Xiao, W.; Loscalzo, J. Metabolic Responses to Reductive Stress. Antioxid. Redox Signal. 2020, 32, 1330–1347. [Google Scholar] [CrossRef] [PubMed]
  59. Maldonado, E.N.; Lemasters, J.J. ATP/ADP ratio, the missed connection between mitochondria and the Warburg effect. Mitochondrion 2014, 19, 78–84. [Google Scholar] [CrossRef] [PubMed]
  60. Jouaville, L.S.; Pinton, P.; Bastianutto, C.; Rutter, G.A.; Rizzuto, R. Regulation of mitochondrial ATP synthesis by calcium: Evidence for a long-term metabolic priming. Proc. Natl. Acad. Sci. USA 1999, 96, 13807–13812. [Google Scholar] [CrossRef]
  61. Jacobson, J.; Duchen, M.R. Interplay between mitochondria and cellular calcium signalling. Mol. Cell. Biochem. 2004, 256, 209–218. [Google Scholar] [CrossRef] [PubMed]
  62. Palmer, C.S.; Osellame, L.D.; Stojanovski, D.; Ryan, M.T. The regulation of mitochondrial morphology: Intricate mechanisms and dynamic machinery. Cell. Signal. 2011, 23, 1534–1545. [Google Scholar] [CrossRef]
  63. O’Leary, B.M.; Asao, S.; Millar, A.H.; Atkin, O.K. Core principles which explain variation in respiration across biological scales. New Phytol. 2018, 222, 670–686. [Google Scholar] [CrossRef] [PubMed]
  64. Coen, P.M.; Jubrias, S.A.; Distefano, G.; Amati, F.; Mackey, D.C.; Glynn, N.W.; Manini, T.M.; Wohlgemuth, S.E.; Leeuwenburgh, C.; Cummings, S.R.; et al. Skeletal Muscle Mitochondrial Energetics Are Associated With Maximal Aerobic Capacity and Walking Speed in Older Adults. J. Gerontol. Ser. A 2012, 68, 447–455. [Google Scholar] [CrossRef]
  65. Berg, S.A.v.D.; Nabben, M.; Bijland, S.; Voshol, P.J.; van Klinken, J.B.; Havekes, L.M.; Romijn, J.A.; Hoeks, J.; Hesselink, M.K.; Schrauwen, P.; et al. High levels of whole-body energy expenditure are associated with a lower coupling of skeletal muscle mitochondria in C57Bl/6 mice. Metabolism 2010, 59, 1612–1618. [Google Scholar] [CrossRef]
  66. Monternier, P.-A.; Marmillot, V.; Rouanet, J.-L.; Roussel, D. Mitochondrial phenotypic flexibility enhances energy savings during winter fast in king penguin chicks. J. Exp. Biol. 2014, 217, 2691–7. [Google Scholar] [CrossRef] [PubMed]
  67. Jacobs, R.A.; Siebenmann, C.; Hug, M.; Toigo, M.; Lundby, C.; Meinild, A.-K. Twenty-eight days at 3454-m altitude diminishes respiratory capacity but enhances efficiency in human skeletal muscle mitochondria. FASEB J. 2012, 26, 5192–5200. [Google Scholar] [CrossRef] [PubMed]
  68. Hulbert, A.J.; Turner, N.; Hinde, J.; Else, P.; Guderley, H. How might you compare mitochondria from different tissues and different species? J. Comp. Physiol. B 2006, 176, 93–105. [Google Scholar] [CrossRef] [PubMed]
  69. Salin, K.; Luquet, E.; Rey, B.; Roussel, D.; Voituron, Y. Alteration of mitochondrial efficiency affects oxidative balance, development and growth in frog (Rana temporaria) tadpoles. J. Exp. Biol. 2012, 215, 863–869. [Google Scholar] [CrossRef] [PubMed]
  70. Brand The efficiency and plasticity of mitochondrial energy transduction. Biochem. Soc. Trans. 2005, 33, 897–904. [CrossRef] [PubMed]
  71. Park, S.-Y.; Gifford, J.R.; Andtbacka, R.H.I.; Trinity, J.D.; Hyngstrom, J.R.; Garten, R.S.; Diakos, N.A.; Ives, S.J.; Dela, F.; Larsen, S.; et al. Cardiac, skeletal, and smooth muscle mitochondrial respiration: Are all mitochondria created equal? Am. J. Physiol. Heart Circ. Physiol. 2014, 307, H346–H352. [Google Scholar] [CrossRef]
  72. Holmström, M.H.; Iglesias-Gutierrez, E.; Zierath, J.R.; Garcia-Roves, P.M. Tissue-specific control of mitochondrial respiration in obesity-related insulin resistance and diabetes. Am. J. Physiol. Metab. 2012, 302, E731–E739. [Google Scholar] [CrossRef] [PubMed]
  73. Salin, K.; Auer, S.K.; Rey, B.; Selman, C.; Metcalfe, N.B. Variation in the link between oxygen consumption and ATP production, and its relevance for animal performance. Proc. R. Soc. B: Boil. Sci. 2015, 282, 20151028. [Google Scholar] [CrossRef]
  74. Burton, T.; Killen, S.S.; Armstrong, J.D.; Metcalfe, N.B. What causes intraspecific variation in resting metabolic rate and what are its ecological consequences? Proc. R. Soc. B Boil. Sci. 2011, 278, 3465–3473. [Google Scholar] [CrossRef] [PubMed]
  75. Padalko, V.I. Uncoupler of Oxidative Phosphorylation Prolongs the Lifespan of Drosophila. Biochemistry 2005, 70, 986–989. [Google Scholar] [CrossRef] [PubMed]
  76. Blount, J.D.; Vitikainen, E.I.K.; Stott, I.; Cant, M.A. Oxidative shielding and the cost of reproduction. Biol. Rev. 2015, 91, 483–497. [Google Scholar] [CrossRef] [PubMed]
  77. Porter, R.K.; Brand, M.D. Body mass dependence of H+ leak in mitochondria and its relevance to metabolic rate. Nature 1993, 362, 628–630. [Google Scholar] [CrossRef]
  78. Wheaton, W.W.; Chandel, N.S.; Chihanga, T.; Ruby, H.N.; Ma, Q.; Bashir, S.; Devarajan, P.; Kennedy, M.A.; Bernard, O.; Jeny, F.; et al. Hypoxia. 2. Hypoxia regulates cellular metabolism. Am. J. Physiol. Physiol. 2011, 300, C385–C393. [Google Scholar] [CrossRef] [PubMed]
  79. Harper, J.A.; Dickinson, K.; Brand, M.D. Mitochondrial uncoupling as a target for drug development for the treatment of obesity. Obes. Rev. 2001, 2, 255–265. [Google Scholar] [CrossRef]
  80. Hinkle, P.C. P/O ratios of mitochondrial oxidative phosphorylation. Biochim. et Biophys. Acta (BBA)-Bioenerg. 2005, 1706, 1–11. [Google Scholar] [CrossRef]
  81. Sommer, A.M.; Pörtner, H.O. Mitochondrial Function in Seasonal Acclimatization versus Latitudinal Adaptation to Cold in the Lugworm Arenicola marina(L.). Physiol. Biochem. Zoöl. 2004, 77, 174–186. [Google Scholar] [CrossRef]
  82. Cannon, B.; Nedergaard, J.; Škop, V.; Liu, N.; Guo, J.; Gavrilova, O.; Reitman, M.L.; Ono-Moore, K.D.; Rutkowsky, J.M.; A Pearson, N.; et al. Brown Adipose Tissue: Function and Physiological Significance. Physiol. Rev. 2004, 84, 277–359. [Google Scholar] [CrossRef]
  83. Teulier, L.; Rouanet, J.-L.; Letexier, D.; Romestaing, C.; Belouze, M.; Rey, B.; Duchamp, C.; Roussel, D. Cold-acclimation-induced non-shivering thermogenesis in birds is associated with upregulation of avian UCP but not with innate uncoupling or altered ATP efficiency. J. Exp. Biol. 2010, 213, 2476–2482. [Google Scholar] [CrossRef]
  84. Meyer, E.H.; Tomaz, T.; Carroll, A.J.; Estavillo, G.; Delannoy, E.; Tanz, S.K.; Small, I.D.; Pogson, B.J.; Millar, A.H. Remodeled Respiration in ndufs4 with Low Phosphorylation Efficiency Suppresses Arabidopsis Germination and Growth and Alters Control of Metabolism at Night. Plant Physiol. 2009, 151, 603–619. [Google Scholar] [CrossRef]
  85. Rasmusson, A.G.; Geisler, D.A.; Møller, I.M. The multiplicity of dehydrogenases in the electron transport chain of plant mitochondria. Mitochondrion 2008, 8, 47–60. [Google Scholar] [CrossRef] [PubMed]
  86. Garvin, M.R.; Thorgaard, G.H.; Narum, S.R. Differential Expression of Genes that Control Respiration Contribute to Thermal Adaptation in Redband Trout ( Oncorhynchus mykiss gairdneri ). Genome Biol. Evol. 2015, 7, 1404–1414. [Google Scholar] [CrossRef] [PubMed]
  87. Bevilacqua, L.; Ramsey, J.J.; Hagopian, K.; Weindruch, R.; Harper, M.-E. Long-term caloric restriction increases UCP3 content but decreases proton leak and reactive oxygen species production in rat skeletal muscle mitochondria. Am. J. Physiol. Metab. 2005, 289, E429–E438. [Google Scholar] [CrossRef] [PubMed]
  88. Yu, L.; Fink, B.D.; Herlein, J.A.; Oltman, C.L.; Lamping, K.G.; Sivitz, W.I. Dietary fat, fatty acid saturation and mitochondrial bioenergetics. J. Bioenerg. Biomembr. 2013, 46, 33–44. [Google Scholar] [CrossRef] [PubMed]
  89. Chávez, M.D.; Tse, H.M. Targeting Mitochondrial-Derived Reactive Oxygen Species in T Cell-Mediated Autoimmune Diseases. Front. Immunol. 2021, 12, 703972. [Google Scholar] [CrossRef] [PubMed]
  90. Schieber, M.; Chandel, N.S. ROS Function in Redox Signaling and Oxidative Stress. Curr. Biol. 2014, 24, R453–R462. [Google Scholar] [CrossRef] [PubMed]
  91. Murphy, M.P. How mitochondria produce reactive oxygen species. Biochem. J. 2009, 417, 1–13. [Google Scholar] [CrossRef] [PubMed]
  92. Korshunov, S.S.; Skulachev, V.P.; Starkov, A.A. High protonic potential actuates a mechanism of production of reactive oxygen species in mitochondria. FEBS Lett. 1997, 416, 15–18. [Google Scholar] [CrossRef] [PubMed]
  93. Keipert, S.; Voigt, A.; Klaus, S. Dietary effects on body composition, glucose metabolism, and longevity are modulated by skeletal muscle mitochondrial uncoupling in mice. Aging Cell 2010, 10, 122–136. [Google Scholar] [CrossRef] [PubMed]
  94. Barja, G. Mitochondrial Oxygen Consumption and Reactive Oxygen Species Production are Independently Modulated: Implications for Aging Studies. Rejuvenation Res. 2007, 10, 215–224. [Google Scholar] [CrossRef]
  95. Salin, K.; Villasevil, E.M.; Anderson, G.J.; Selman, C.; Chinopoulos, C.; Metcalfe, N.B. The RCR and ATP/O Indices Can Give Contradictory Messages about Mitochondrial Efficiency. Integr. Comp. Biol. 2018, 58, 486–494. [Google Scholar] [CrossRef] [PubMed]
  96. Toyomizu, M.; Kikusato, M.; Kawabata, Y.; Azad, A.K.; Inui, E.; Amo, T. Meat-type chickens have a higher efficiency of mitochondrial oxidative phosphorylation than laying-type chickens. Comp. Biochem. Physiol. Part A Mol. Integr. Physiol. 2011, 159, 75–81. [Google Scholar] [CrossRef] [PubMed]
  97. Tinsley, N.; Iqbal, M.; Pumford, N.R.; Lassiter, K.; Ojano-Dirain, C.; Wing, T.; Bottje, W. Investigation of mitochondrial protein expression and oxidation in heart muscle in low and high feed efficient male broilers in a single genetic line. Poult. Sci. 2010, 89, 349–352. [Google Scholar] [CrossRef]
  98. Figueroa, C.M.; Lunn, J.E. A Tale of Two Sugars: Trehalose 6-Phosphate and Sucrose. Plant Physiol. 2016, 172, 7–27. [Google Scholar] [CrossRef] [PubMed]
  99. Millar, A.H.; Atkin, O.K.; Menz, R.I.; Henry, B.; Farquhar, G.; Day, D.A. Analysis of Respiratory Chain Regulation in Roots of Soybean Seedlings1. Plant Physiol. 1998, 117, 1083–1093. [Google Scholar] [CrossRef] [PubMed]
  100. Millar, A.H.; Whelan, J.; Soole, K.L.; Day, D.A. Organization and Regulation of Mitochondrial Respiration in Plants. Annu. Rev. Plant Biol. 2011, 62, 79–104. [Google Scholar] [CrossRef] [PubMed]
  101. Amthor, J.S. The McCree–de Wit–Penning de Vries–Thornley respiration paradigms: 30 years later. Ann. Bot. 2000, 86, 1–20. [Google Scholar] [CrossRef]
  102. Lucia, U. Bioengineering thermodynamics of biological cells. Theor. Biol. Med Model. 2015, 12, 29. [Google Scholar] [CrossRef]
  103. Nunn, A.V.; Guy, G.W.; Bell, J.D. The quantum mitochondrion and optimal health. Biochem. Soc. Trans. 2016, 44, 1101–1110. [Google Scholar] [CrossRef]
  104. Thomsen, K. Timelessness Strictly inside the Quantum Realm. Entropy 2021, 23, 772. [Google Scholar] [CrossRef] [PubMed]
  105. Bera, M.N.; Acín, A.; Kuś, M.; Mitchell, M.W.; Lewenstein, M. Randomness in quantum mechanics: Philosophy, physics and technology. Rep. Prog. Phys. 2017, 80, 124001. [Google Scholar] [CrossRef] [PubMed]
  106. Kennett, E.; Kuchel, P. Redox Reactions and Electron Transfer Across the Red Cell Membrane. IUBMB Life 2003, 55, 375–385. [Google Scholar] [CrossRef] [PubMed]
  107. Geigenberger, P.; Fernie, A.R. Metabolic Control of Redox and Redox Control of Metabolism in Plants. Antioxid. Redox Signal. 2014, 21, 1389–1421. [Google Scholar] [CrossRef]
  108. Xiao, W.; Wang, R.-S.; Handy, D.E.; Loscalzo, J.; Md; Zhao, Y.; Zhang, Z.; Zou, Y.; Yang, Y.; Fessel, J.P.; et al. NAD(H) and NADP(H) Redox Couples and Cellular Energy Metabolism. Antioxid. Redox Signal. 2018, 28, 251–272. [Google Scholar] [CrossRef]
  109. Mailloux, R.J.; Jin, X.; Willmore, W.G. Redox regulation of mitochondrial function with emphasis on cysteine oxidation reactions. Redox Biol. 2013, 2, 123–139. [Google Scholar] [CrossRef]
  110. Sutcliffe, M.J.; Scrutton, N.S. Enzymology takes a quantum leap forward. Philos. Trans. R. Soc. A Math. Phys. Eng. Sci. 2000, 358, 367–386. [Google Scholar] [CrossRef]
  111. Hernansanz-Agustín, P.; Enríquez, J.A. Generation of Reactive Oxygen Species by Mitochondria. Antioxidants 2021, 10, 415. [Google Scholar] [CrossRef]
  112. Xin, H.; Sim, W.J.; Namgung, B.; Choi, Y.; Li, B.; Lee, L.P. Quantum biological tunnel junction for electron transfer imaging in live cells. Nat. Commun. 2019, 10, 1–11. [Google Scholar] [CrossRef]
  113. Tyburski, R.; Liu, T.; Glover, S.D.; Hammarström, L. Proton-Coupled Electron Transfer Guidelines, Fair and Square. J. Am. Chem. Soc. 2021, 143, 560–576. [Google Scholar] [CrossRef]
  114. Usselman, R.J.; Chavarriaga, C.; Castello, P.R.; Procopio, M.; Ritz, T.; Dratz, E.A.; Singel, D.J.; Martino, C.F. The Quantum Biology of Reactive Oxygen Species Partitioning Impacts Cellular Bioenergetics. Sci. Rep. 2016, 6, 38543. [Google Scholar] [CrossRef] [PubMed]
  115. Canfield, J.; Belford, R.; Debrunner, P.; Schulten, K. A perturbation theory treatment of oscillating magnetic fields in the radical pair mechanism. Chem. Phys. 1994, 182, 1–18. [Google Scholar] [CrossRef]
  116. Imlay, J.A. The molecular mechanisms and physiological consequences of oxidative stress: Lessons from a model bacterium. Nat. Rev. Genet. 2013, 11, 443–454. [Google Scholar] [CrossRef] [PubMed]
  117. Usselman, R.J.; Hill, I.; Singel, D.J.; Martino, C.F. Spin Biochemistry Modulates Reactive Oxygen Species (ROS) Production by Radio Frequency Magnetic Fields. PLoS ONE 2014, 9, e93065. [Google Scholar] [CrossRef] [PubMed]
  118. Hogben, H.J.; Efimova, O.; Wagner-Rundell, N.; Timmel, C.R.; Hore, P. Possible involvement of superoxide and dioxygen with cryptochrome in avian magnetoreception: Origin of Zeeman resonances observed by in vivo EPR spectroscopy. Chem. Phys. Lett. 2009, 480, 118–122. [Google Scholar] [CrossRef]
  119. Brand, K.A.; Hermfisse, U. Aerobic glycolysis by proliferating cells: A protective strategy against reactive oxygen species. FASEB J. 1997, 11, 388–395. [Google Scholar] [CrossRef]
  120. Liemburg-Apers, D.C.; Willems, P.H.G.M.; Koopman, W.J.H.; Grefte, S. Interactions between mitochondrial reactive oxygen species and cellular glucose metabolism. Arch. Toxicol. 2015, 89, 1209–1226. [Google Scholar] [CrossRef]
  121. Willems, P.H.; Rossignol, R.; Dieteren, C.E.; Murphy, M.P.; Koopman, W.J. Redox Homeostasis and Mitochondrial Dynamics. Cell Metab. 2015, 22, 207–218. [Google Scholar] [CrossRef]
  122. Ji, L.L.; Kang, C.; Zhang, Y. Exercise-induced hormesis and skeletal muscle health. Free. Radic. Biol. Med. 2016, 98, 113–122. [Google Scholar] [CrossRef]
  123. Marais, A.; Adams, B.; Ringsmuth, A.K.; Ferretti, M.; Gruber, J.M.; Hendrikx, R.; Schuld, M.; Smith, S.L.; Sinayskiy, I.; Krüger, T.P.J.; et al. The future of quantum biology. J. R. Soc. Interface 2018, 15, 20180640. [Google Scholar] [CrossRef]
  124. Brack, T.; Jubran, A.; Tobin, M.J. Dyspnea and Decreased Variability of Breathing in Patients with Restrictive Lung Disease. Am. J. Respir. Crit. Care Med. 2002, 165, 1260–1264. [Google Scholar] [CrossRef] [PubMed]
  125. Loveridge, B.; West, P.; Anthonisen, N.R.; Kryger, M.H. Breathing patterns in patients with chronic obstructive pulmonary disease. Am. Rev. Respir. Dis. 1984, 130. [Google Scholar] [CrossRef]
  126. Schechtman, V.L.; Lee, M.Y.; Wilson, A.J.; Harper, R.M. Dynamics of Respiratory Patterning in Normal Infants and Infants Who Subsequently Died of the Sudden Infant Death Syndrome. Pediatr. Res. 1996, 40, 571–577. [Google Scholar] [CrossRef] [PubMed]
  127. Thamrin, C.; Frey, U.; Kaminsky, D.A.; Reddel, H.K.; Seely, A.J.E.; Suki, B.; Sterk, P.J. Systems Biology and Clinical Practice in Respiratory Medicine. The Twain Shall Meet. Am. J. Respir. Crit. Care Med. 2016, 194, 1053–1061. [Google Scholar] [CrossRef]
  128. Yentes, J.M.; Fallahtafti, F.; Denton, W.; Rennard, S.I. COPD Patients Have a Restricted Breathing Pattern That Persists with Increased Metabolic Demands. COPD: J. Chronic Obstr. Pulm. Dis. 2020, 17, 245–252. [Google Scholar] [CrossRef]
  129. Usemann, J.; Suter, A.; Zannin, E.; Proietti, E.; Fouzas, S.; Schulzke, S.; Latzin, P.; Frey, U.; Fuchs, O.; Korten, I.; et al. Variability of Tidal Breathing Parameters in Preterm Infants and Associations with Respiratory Morbidity during Infancy: A Cohort Study. J. Pediatr. 2019, 205, 61–69.e1. [Google Scholar] [CrossRef]
  130. Hmeidi, H.; Motamedi-Fakhr, S.; Chadwick, E.; Gilchrist, F.J.; Lenney, W.; Iles, R.; Wilson, R.C.; Alexander, J. Tidal breathing parameters measured using structured light plethysmography in healthy children and those with asthma before and after bronchodilator. Physiol. Rep. 2017, 5, e13168. [Google Scholar] [CrossRef]
  131. Thamrin, C.; Taylor, D.R.; Jones, S.L.; Suki, B.; Frey, U. Variability of lung function predicts loss of asthma control following withdrawal of inhaled corticosteroid treatment. Thorax 2010, 65, 403–408. [Google Scholar] [CrossRef]
  132. Thamrin, C.; Stern, G.; Strippoli, M.-P.F.; Kuehni, C.E.; Suki, B.; Taylor, D.R.; Frey, U. Fluctuation analysis of lung function as a predictor of long-term response to 2-agonists. Eur. Respir. J. 2008, 33, 486–493. [Google Scholar] [CrossRef]
  133. Frey, U.; Brodbeck, T.; Majumdar, A.; Taylor, D.R.; Town, G.I.; Silverman, M.; Suki, B. Risk of severe asthma episodes predicted from fluctuation analysis of airway function. Nature 2005, 438, 667–670. [Google Scholar] [CrossRef]
  134. Oostveen, E.; MacLeod, D.; Lorino, H.; Farre, R.; Hantos, Z.; Desager, K.; Marchal, F. The forced oscillation technique in clinical practice: Methodology, recommendations and future developments. Eur. Respir. J. 2003, 22, 1026–1041. [Google Scholar] [CrossRef] [PubMed]
  135. Gobbi, A.; Dellacá, R.L.; King, G.; Thamrin, C. Toward Predicting Individual Risk in Asthma Using Daily Home Monitoring of Resistance. Am. J. Respir. Crit. Care Med. 2017, 195, 265–267. [Google Scholar] [CrossRef]
  136. Thamrin, C.; Zindel, J.; Nydegger, R.; Reddel, H.K.; Chanez, P.; Wenzel, S.E.; FitzPatrick, S.; Watt, R.A.; Suki, B.; Frey, U. Predicting future risk of asthma exacerbations using individual conditional probabilities. J. Allergy Clin. Immunol. 2011, 127, 1494–1502.e3. [Google Scholar] [CrossRef] [PubMed]
  137. Timmins, S.C.; Coatsworth, N.; Palnitkar, G.; Thamrin, C.; Farrow, C.E.; Schoeffel, R.E.; Berend, N.; Diba, C.; Salome, C.M.; King, G.G. Day-to-day variability of oscillatory impedance and spirometry in asthma and COPD. Respir. Physiol. Neurobiol. 2013, 185, 416–424. [Google Scholar] [CrossRef]
  138. Zimmermann, S.C.; Huvanandana, J.; Nguyen, C.D.; Bertolin, A.; Watts, J.C.; Gobbi, A.; Farah, C.S.; Peters, M.J.; Dellacà, R.L.; King, G.G.; et al. Day-to-day variability of forced oscillatory mechanics for early detection of acute exacerbations in COPD. Eur. Respir. J. 2020, 56, 1901739. [Google Scholar] [CrossRef]
  139. Veit, T.; Barnikel, M.; Crispin, A.; Kneidinger, N.; Ceelen, F.; Arnold, P.; Munker, D.; Schmitzer, M.; Barton, J.; Schiopu, S.; et al. Variability of forced vital capacity in progressive interstitial lung disease: A prospective observational study. Respir. Res. 2020, 21, 1–10. [Google Scholar] [CrossRef] [PubMed]
  140. Askanazi, J.; Silverberg, P.A.; Hyman, A.I.; Rosenbaum, S.H.; Foster, R.; Kinney, J.M. Patterns of ventilation in postoperative and acutely ill patients. Crit. Care Med. 1979, 7, 41–46. [Google Scholar] [CrossRef] [PubMed]
  141. Preas, H.L.; Jubran, A.; Vandivier, R.W.; Reda, D.; Godin, P.; Banks, S.M.; Tobin, M.J.; Suffredini, F. Effect of endotoxin on ventilation and breath variability: Role of cyclooxygenase pathway. Am. J. Respir. Crit. Care Med. 2001, 164, 620–626. [Google Scholar] [CrossRef]
  142. Bien, M.-Y.; Hseu, S.-S.; Yien, H.-W.; Kuo, B.I.-T.; Lin, Y.-T.; Wang, J.-H.; Kou, Y.R. Breathing pattern variability: A weaning predictor in postoperative patients recovering from systemic inflammatory response syndrome. Intensiv. Care Med. 2004, 30, 241–247. [Google Scholar] [CrossRef] [PubMed]
  143. Brigham, K.; Meyrick, B. Endotoxin and lung injury. Am. Rev. Respir. Dis. 1986, 133, 913–927. [Google Scholar]
  144. Ilan, Y. Making use of noise in biological systems. Prog. Biophys. Mol. Biol. 2023, 178, 83–90. [Google Scholar] [CrossRef] [PubMed]
  145. Haney, S.; Hancox, R.J. Rapid onset of tolerance to beta-agonist bronchodilation. Respir. Med. 2005, 99, 566–571. [Google Scholar] [CrossRef] [PubMed]
  146. Haney, S.; Hancox, R.J. Overcoming beta-agonist tolerance: High dose salbutamol and ipratropium bromide. Two randomised controlled trials. Respir. Res. 2007, 8, 19. [Google Scholar] [CrossRef]
  147. Wraight, J.; Hancox, R.; Herbison, G.; Cowan, J.; Flannery, E.; Taylor, D. Bronchodilator tolerance: The impact of increasing bronchoconstriction. Eur. Respir. J. 2003, 21, 810–815. [Google Scholar] [CrossRef] [PubMed]
  148. Kessler, A.; Weksler-Zangen, S.; Ilan, Y. Role of the Immune System and the Circadian Rhythm in the Pathogenesis of Chronic Pancreatitis. Pancreas 2020, 49, 1024–1032. [Google Scholar] [CrossRef] [PubMed]
  149. Ishay, Y.; Kolben, Y.; Kessler, A.; Ilan, Y. Role of circadian rhythm and autonomic nervous system in liver function: A hypothetical basis for improving the management of hepatic encephalopathy. Am. J. Physiol. Liver Physiol. 2021, 321, G400–G412. [Google Scholar] [CrossRef] [PubMed]
  150. Kolben, Y.; Weksler-Zangen, S.; Ilan, Y. Adropin as a potential mediator of the metabolic system-autonomic nervous system-chronobiology axis: Implementing a personalized signature-based platform for chronotherapy. Obes. Rev. 2020, 22, e13108. [Google Scholar] [CrossRef]
  151. Kenig, A.; Kolben, Y.; Asleh, R.; Amir, O.; Ilan, Y. Improving Diuretic Response in Heart Failure by Implementing a Patient-Tailored Variability and Chronotherapy-Guided Algorithm. Front. Cardiovasc. Med. 2021, 8, 695547. [Google Scholar] [CrossRef]
  152. Azmanov, H.; Ross, E.L.; Ilan, Y. Establishment of an Individualized Chronotherapy, Autonomic Nervous System, and Variability-Based Dynamic Platform for Overcoming the Loss of Response to Analgesics. Pain Physician 2021, 24, 243–252. [Google Scholar] [PubMed]
  153. Potruch, A.; Khoury, S.T.; Ilan, Y. The role of chronobiology in drug-resistance epilepsy: The potential use of a variability and chronotherapy-based individualized platform for improving the response to anti-seizure drugs. Seizure 2020, 80, 201–211. [Google Scholar] [CrossRef] [PubMed]
  154. Isahy, Y.; Ilan, Y. Improving the long-term response to antidepressants by establishing an individualized platform based on variability and chronotherapy. Int. J. Clin. Pharmacol. Ther. 2021, 59, 768–774. [Google Scholar] [CrossRef]
  155. Khoury, T.; Ilan, Y. Introducing Patterns of Variability for Overcoming Compensatory Adaptation of the Immune System to Immunomodulatory Agents: A Novel Method for Improving Clinical Response to Anti-TNF Therapies. Front. Immunol. 2019, 10, 2726. [Google Scholar] [CrossRef] [PubMed]
  156. Khoury, T.; Ilan, Y. Platform introducing individually tailored variability in nerve stimulations and dietary regimen to prevent weight regain following weight loss in patients with obesity. Obes. Res. Clin. Pract. 2021, 15, 114–123. [Google Scholar] [CrossRef]
  157. Kenig, A.; Ilan, Y. A Personalized Signature and Chronotherapy-Based Platform for Improving the Efficacy of Sepsis Treatment. Front. Physiol. 2019, 10, 1542. [Google Scholar] [CrossRef] [PubMed]
  158. Ilan, Y. Why targeting the microbiome is not so successful: Can randomness overcome the adaptation that occurs following gut manipulation? Clin. Exp. Gastroenterol. 2019, 12, 209–217. [Google Scholar] [CrossRef]
  159. Gelman, R.; Bayatra, A.; Kessler, A.; Schwartz, A.; Ilan, Y. Targeting SARS-CoV-2 receptors as a means for reducing infectivity and improving antiviral and immune response: An algorithm-based method for overcoming resistance to antiviral agents. Emerg. Microbes Infect. 2020, 9, 1397–1406. [Google Scholar] [CrossRef] [PubMed]
  160. Ishay, Y.; Potruch, A.; Schwartz, A.; Berg, M.; Jamil, K.; Agus, S.; Ilan, Y. A digital health platform for assisting the diagnosis and monitoring of COVID-19 progression: An adjuvant approach for augmenting the antiviral response and mitigating the immune-mediated target organ damage. BioMedicine 2021, 143, 112228. [Google Scholar] [CrossRef] [PubMed]
  161. Ilan, Y.; Spigelman, Z. Establishing patient-tailored variability-based paradigms for anti-cancer therapy: Using the inherent trajectories which underlie cancer for overcoming drug resistance. Cancer Treat. Res. Commun. 2020, 25, 100240. [Google Scholar] [CrossRef] [PubMed]
  162. Hurvitz, N.; Azmanov, H.; Kesler, A.; Ilan, Y. Establishing a second-generation artificial intelligence-based system for improving diagnosis, treatment, and monitoring of patients with rare diseases. Eur. J. Hum. Genet. 2021, 29, 1485–1490. [Google Scholar] [CrossRef] [PubMed]
  163. Ilan, Y. Digital Medical Cannabis as Market Differentiator: Second-Generation Artificial Intelligence Systems to Improve Response. Front. Med. 2022, 8, 788777. [Google Scholar] [CrossRef] [PubMed]
  164. Gelman, R.; Berg, M.; Ilan, Y. A Subject-Tailored Variability-Based Platform for Overcoming the Plateau Effect in Sports Training: A Narrative Review. Int. J. Environ. Res. Public Heal. 2022, 19, 1722. [Google Scholar] [CrossRef] [PubMed]
  165. Azmanov, H.; Bayatra, A.; Ilan, Y. Digital Analgesic Comprising a Second-Generation Digital Health System: Increasing Effectiveness by Optimizing the Dosing and Minimizing Side Effects. J. Pain Res. 2022, 15, 1051–1060. [Google Scholar] [CrossRef] [PubMed]
  166. Kolben, Y.; Azmanov, H.; Gelman, R.; Dror, D.; Ilan, Y. Using chronobiology-based second-generation artificial intelligence digital system for overcoming antimicrobial drug resistance in chronic infections. Ann. Med. 2023, 55, 311–318. [Google Scholar] [CrossRef] [PubMed]
  167. Hurvitz, N.; Elkhateeb, N.; Sigawi, T.; Rinsky-Halivni, L.; Ilan, Y. Improving the effectiveness of anti-aging modalities by using the constrained disorder principle-based management algorithms. Front. Aging 2022, 3, 1044038. [Google Scholar] [CrossRef]
  168. Ilan, Y. Next-Generation Personalized Medicine: Implementation of Variability Patterns for Overcoming Drug Resistance in Chronic Diseases. J. Pers. Med. 2022, 12, 1303. [Google Scholar] [CrossRef] [PubMed]
  169. Hurvitz, N.; Ilan, Y. The Constrained-Disorder Principle Assists in Overcoming Significant Challenges in Digital Health: Moving from “Nice to Have” to Mandatory Systems. Clin. Pract. 2023, 13, 994–1014. [Google Scholar] [CrossRef] [PubMed]
  170. Ilan, Y. Constrained disorder principle-based variability is fundamental for biological processes: Beyond biological relativity and physiological regulatory networks. Prog. Biophys. Mol. Biol. 2023, 180–181, 37–48. [Google Scholar] [CrossRef]
  171. Sigawi, T.; Lehmann, H.; Hurvitz, N.; Ilan, Y. Constrained disorder principle-based second-generation algorithms implement quantified variability signatures to improve the function of complex systems. J. Bioinform. Syst. Biol. 2023, 06, 82–89. [Google Scholar] [CrossRef]
  172. Gelman, R.; Hurvitz, N.; Nesserat, R.; Kolben, Y.; Nachman, D.; Jamil, K.; Agus, S.; Asleh, R.; Amir, O.; Berg, M.; et al. A second-generation artificial intelligence-based therapeutic regimen improves diuretic resistance in heart failure: Results of a feasibility open-labeled clinical trial. BioMedicine 2023, 161, 114334. [Google Scholar] [CrossRef]
  173. Suki, B.; Alencar, A.M.; Sujeer, M.K.; Lutchen, K.R.; Collins, J.J.; Andrade, J.S.; Ingenito, E.P.; Zapperi, S.; Stanley, H.E. Life-support system benefits from noise. Nature 1998, 393, 127–128. [Google Scholar] [CrossRef]
  174. Brewster, J.F.; Graham, M.R.; Mutch, W.A.C. Convexity, Jensen’s inequality and benefits of noisy mechanical ventilation. J. R. Soc. Interface 2005, 2, 393–396. [Google Scholar] [CrossRef] [PubMed]
  175. de Abreu, M.G.; Spieth, P.M.; Pelosi, P.; Carvalho, A.R.; Walter, C.; Schreiber-Ferstl, A.; Aikele, P.; Neykova, B.; Hübler, M.; Koch, T. Noisy pressure support ventilation: A pilot study on a new assisted ventilation mode in experimental lung injury*. Crit. Care Med. 2008, 36, 818–827. [Google Scholar] [CrossRef] [PubMed]
  176. Lefevre, G.R.; E Kowalski, S.; Girling, L.G.; Thiessen, D.B.; A Mutch, W. Improved arterial oxygenation after oleic acid lung injury in the pig using a computer-controlled mechanical ventilator. Am. J. Respir. Crit. Care Med. 1996, 154, 1567–1572. [Google Scholar] [CrossRef]
  177. Arold, S.P.; Suki, B.; Alencar, A.M.; Lutchen, K.R.; Ingenito, E.P. Variable ventilation induces endogenous surfactant release in normal guinea pigs. Am. J. Physiol. Cell. Mol. Physiol. 2003, 285, L370–L375. [Google Scholar] [CrossRef]
  178. Graham, M.R.; Goertzen, A.L.; Girling, L.G.; Friedman, T.; Pauls, R.J.; Dickson, T.; Espenell, A.E.G.; Mutch, W.A.C. Quantitative computed tomography in porcine lung injury with variable versus conventional ventilation: Recruitment and surfactant replacement*. Crit. Care Med. 2011, 39, 1721–1730. [Google Scholar] [CrossRef]
  179. Bellardine, C.L.; Hoffman, A.M.; Tsai, L.; Ingenito, E.P.; Arold, S.P.; Lutchen, K.R.; Suki, B. Comparison of variable and conventional ventilation in a sheep saline lavage lung injury model*. Crit. Care Med. 2006, 34, 439–445. [Google Scholar] [CrossRef]
  180. Mutch, W.A.C.; Harms, S.; Lefevre, G.R.; Graham, M.R.; Girling, L.G.; Kowalski, S.E. Biologically variable ventilation increases arterial oxygenation over that seen with positive end-expiratory pressure alone in a porcine model of acute respiratory distress syndrome. Crit. Care Med. 2000, 28, 2457–2464. [Google Scholar] [CrossRef]
  181. Arold, S.P.; Mora, R.; Lutchen, K.R.; Ingenito, E.P.; Suki, B. Variable Tidal Volume Ventilation Improves Lung Mechanics and Gas Exchange in a Rodent Model of Acute Lung Injury. Am. J. Respir. Crit. Care Med. 2002, 165, 366–371. [Google Scholar] [CrossRef] [PubMed]
  182. Spieth, P.M.; Carvalho, A.R.; Pelosi, P.; Hoehn, C.; Meissner, C.; Kasper, M.; Hübler, M.; von Neindorff, M.; Dassow, C.; Barrenschee, M.; et al. Variable Tidal Volumes Improve Lung Protective Ventilation Strategies in Experimental Lung Injury. Am. J. Respir. Crit. Care Med. 2009, 179, 684–693. [Google Scholar] [CrossRef] [PubMed]
  183. Graham, M.R.; Gulati, H.; Kha, L.; Girling, L.G.; Goertzen, A.; Mutch, W.A.C. Resolution of pulmonary edema with variable mechanical ventilation in a porcine model of acute lung injury. Can. J. Anaesth. 2011, 58, 740–750. [Google Scholar] [CrossRef] [PubMed]
  184. Boker, A.; Haberman, C.J.; Girling, L.; Guzman, R.P.; Louridas, G.; Tanner, J.R.; Cheang, M.; Maycher, B.W.; Bell, D.D.; Doak, G.J. Variable Ventilation Improves Perioperative Lung Function in Patients Undergoing Abdominal Aortic Aneurysmectomy. Anesthesiology 2004, 100, 608–616. [Google Scholar] [CrossRef]
  185. Mutch, W.A.C.; Buchman, T.G.; Girling, L.G.; Walker, E.K.-Y.; McManus, B.M.; Graham, M.R. Biologically variable ventilation improves gas exchange and respiratory mechanics in a model of severe bronchospasm*. Crit. Care Med. 2007, 35, 1749–1755. [Google Scholar] [CrossRef] [PubMed]
  186. Epstein, S.K. Extubation failure: An outcome to be avoided. Crit. Care 2004, 8, 310–2. [Google Scholar] [CrossRef] [PubMed]
  187. Krasinkiewicz, J.M.; Friedman, M.L.; E Slaven, J.; Lutfi, R.; Abu-Sultaneh, S.; Tori, A.J. Extubation Readiness Practices and Barriers to Extubation in Pediatric Subjects. Respir. Care 2020, 66, 582–590. [Google Scholar] [CrossRef] [PubMed]
  188. Pan, Q.; Zhang, H.; Jiang, M.; Ning, G.; Fang, L.; Ge, H. Comprehensive breathing variability indices enhance the prediction of extubation failure in patients on mechanical ventilation. Comput. Biol. Med. 2023, 153, 106459. [Google Scholar] [CrossRef] [PubMed]
  189. Wysocki, M.; Cracco, C.; Teixeira, A.; Mercat, A.; Diehl, J.-L.; Lefort, Y.; Derenne, J.-P.; Similowski, T. Reduced breathing variability as a predictor of unsuccessful patient separation from mechanical ventilation*. Crit. Care Med. 2006, 34, 2076–2083. [Google Scholar] [CrossRef]
  190. Seely, A.J.; Bravi, A.; Herry, C.; Green, G.; Longtin, A.; Ramsay, T.; Fergusson, D.; McIntyre, L.; Kubelik, D.; E Maziak, D.; et al. Do heart and respiratory rate variability improve prediction of extubation outcomes in critically ill patients? Crit. Care 2014, 18, R65. [Google Scholar] [CrossRef]
  191. I Naik, B.; Lynch, C.; Durbin, C.G. Variability in Mechanical Ventilation: What’s All the Noise About? Respir. Care 2015, 60, 1203–1210. [Google Scholar] [CrossRef]
  192. Mauri, T.; Foti, G.; Fornari, C.; Grasselli, G.; Pinciroli, R.; Lovisari, F.; Tubiolo, D.; Volta, C.A.; Spadaro, S.; Rona, R.; et al. Sigh in Patients With Acute Hypoxemic Respiratory Failure and ARDS. Chest 2021, 159, 1426–1436. [Google Scholar] [CrossRef]
  193. Mauri, T.; for the Protection Study Group; Foti, G.; Fornari, C.; Constantin, J.-M.; Guerin, C.; Pelosi, P.; Ranieri, M.; Conti, S.; Tubiolo, D.; et al. Pressure support ventilation + sigh in acute hypoxemic respiratory failure patients: Study protocol for a pilot randomized controlled trial, the PROTECTION trial. Trials 2018, 19, 1–10. [Google Scholar] [CrossRef]
  194. El-Khatib, M.; Jamaleddine, G.; Soubra, R.; Muallem, M. Pattern of spontaneous breathing: Potential marker for weaning outcome. Intensive Care Med. 2001, 27, 52–58. [Google Scholar] [CrossRef] [PubMed]
  195. Zein, H.; Baratloo, A.; Negida, A.; Safari, S. Ventilator Weaning and Spontaneous Breathing Trials; an Educational Review. Emergency 2016, 4, 65–71. [Google Scholar] [PubMed]
  196. Knox, K.E.; Nava-Guerra, L.; Hotz, J.C.R.; Newth, C.J.L.; Khoo, M.C.K.; Khemani, R.G. High Breath-by-Breath Variability Is Associated With Extubation Failure in Children. Crit. Care Med. 2020, 48, 1165–1174. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Overcoming tolerance to bronchodilators by implementing variability-based therapeutic regimens: The figure summarizes the design and function of a CDP-based second-generation AI system for medical treatment, specifically in the context of lung conditions. The system is based on three levels of operation: (A) The AI provides randomized medication regimens within a pre-defined range, aimed to overcome treatment tolerance. (B) The AI collects data on the output and uses clinically meaningful endpoints as inputs to personalize the output, which alters randomization to reach a better outcome. (C) The AI receives inputs from the patient’s quantified variability signatures and uses large datasets from big-data resources to further personalize the treatment for the individual patient. CDP, constrained disorder principle. AI, artificial intelligence. Created with Biorender.com.
Figure 1. Overcoming tolerance to bronchodilators by implementing variability-based therapeutic regimens: The figure summarizes the design and function of a CDP-based second-generation AI system for medical treatment, specifically in the context of lung conditions. The system is based on three levels of operation: (A) The AI provides randomized medication regimens within a pre-defined range, aimed to overcome treatment tolerance. (B) The AI collects data on the output and uses clinically meaningful endpoints as inputs to personalize the output, which alters randomization to reach a better outcome. (C) The AI receives inputs from the patient’s quantified variability signatures and uses large datasets from big-data resources to further personalize the treatment for the individual patient. CDP, constrained disorder principle. AI, artificial intelligence. Created with Biorender.com.
Arm 91 00028 g001
Figure 2. Inserting variability into mechanical ventilation to improve the efficacy of ventilation and ease the weaning process: Breathing variability serves as a predictor for successful extubation and improves the efficacy of mechanical ventilation. (A) Variability in breathing patterns, as measured by indices such as the CV of tidal volume/inspiratory time and CV of Inspiratory time/respiratory period, is an important predictor of extubation success or failure. (B) Variable ventilation was shown to be beneficial in animal models and in human patients, increasing arterial oxygenation, lowering shunt fraction and increasing pulmonary compliance. CV, coefficient of variability. Created with Biorender.com.
Figure 2. Inserting variability into mechanical ventilation to improve the efficacy of ventilation and ease the weaning process: Breathing variability serves as a predictor for successful extubation and improves the efficacy of mechanical ventilation. (A) Variability in breathing patterns, as measured by indices such as the CV of tidal volume/inspiratory time and CV of Inspiratory time/respiratory period, is an important predictor of extubation success or failure. (B) Variable ventilation was shown to be beneficial in animal models and in human patients, increasing arterial oxygenation, lowering shunt fraction and increasing pulmonary compliance. CV, coefficient of variability. Created with Biorender.com.
Arm 91 00028 g002
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Adar, O.; Hollander, A.; Ilan, Y. The Constrained Disorder Principle Accounts for the Variability That Characterizes Breathing: A Method for Treating Chronic Respiratory Diseases and Improving Mechanical Ventilation. Adv. Respir. Med. 2023, 91, 350-367. https://doi.org/10.3390/arm91050028

AMA Style

Adar O, Hollander A, Ilan Y. The Constrained Disorder Principle Accounts for the Variability That Characterizes Breathing: A Method for Treating Chronic Respiratory Diseases and Improving Mechanical Ventilation. Advances in Respiratory Medicine. 2023; 91(5):350-367. https://doi.org/10.3390/arm91050028

Chicago/Turabian Style

Adar, Ofek, Adi Hollander, and Yaron Ilan. 2023. "The Constrained Disorder Principle Accounts for the Variability That Characterizes Breathing: A Method for Treating Chronic Respiratory Diseases and Improving Mechanical Ventilation" Advances in Respiratory Medicine 91, no. 5: 350-367. https://doi.org/10.3390/arm91050028

Article Metrics

Back to TopTop