1. Introduction
Biological information processing has undergone remarkable evolutionary transitions, from primitive molecular recognition systems to the complex neural networks underlying consciousness, yet the physical mechanisms underlying this progression remain incompletely understood (Kandel et al., 2012; McFadden, 2020; Hunt & Jones, 2023). While traditionally conceptualized through distinct mechanistic categories including chemical signaling, bioelectric transmission, and synaptic computation, mounting evidence suggests these processes may be unified under electromagnetic field theory (Ruffini et al., 2020; Anastassiou et al., 2011). Within the Standard Model of particle physics, all molecular interactions in biological systems operate through electromagnetic forces, as chemical bonds represent electromagnetic interactions between charged particles, conformational changes involve electromagnetic field-driven bond reorganization, and cellular signaling proceeds through electromagnetic charge transport and field effects (Weinberg, 1995; Lane, 2015; McFadden & Al-Khalili, 2018). This theoretical framework suggests that apparent mechanistic diversity in biological information processing may reflect different organizational scales of unified electromagnetic dynamics rather than fundamentally distinct physical processes.
Recent discoveries in ephaptic coupling have demonstrated that electromagnetic fields in neural tissue can transmit information at speeds approaching 47 km/s, nearly 500 times faster than the fastest myelinated axons and 5,000 times faster than unmyelinated transmission (Anastassiou et al., 2011; Pinotsis & Miller, 2023; Chiang et al., 2019). Revolutionary new evidence demonstrates that electromagnetic fields can entrain neural spike timing at thresholds as low as 0.74 mV/mm, establishing direct causal field-to-neuron communication that challenges traditional assumptions about neural information flow (Anastassiou et al., 2011; Francis et al., 2003). When this linear speed advantage is considered volumetrically across three-dimensional neural tissue, the theoretical information processing capacity scales as the cube of the linear advantage, yielding approximately 125 billion times greater information density potential for electromagnetic field-based computation compared to neural spike transmission (Ruffini et al., 2020; Hunt & Jones, 2023). These findings challenge conventional models emphasizing chemical synaptic transmission and suggest that electromagnetic field-based processing may be fundamental to neural computation, particularly for functions requiring rapid integration across distributed brain regions (Hunt, 2024; McFadden, 2020; Freeman, 2004). Parallel developments in quantum biology have revealed coherent electromagnetic effects persisting in biological systems despite thermal noise, including quantum coherence in photosynthetic light harvesting (Engel et al., 2007; Collini et al., 2010), avian magnetoreception (Ritz et al., 2000; Mouritsen & Ritz, 2005), and potentially neural microtubule networks (Hameroff & Penrose, 2014; Bandyopadhyay, 2019), indicating that quantum electromagnetic processes may be evolutionarily conserved across diverse biological systems.
Mae-Wan Ho's pioneering theoretical work on biological coherence proposed that living systems exhibit "quantum jazz," characterized by coordinated electromagnetic oscillations that enable near-instantaneous information transfer through structured water networks exhibiting liquid crystalline properties (Ho, 1998, 2008; Ho & Knight, 1998). This framework, when combined with Andrew Parker's "Light Switch" hypothesis linking Cambrian evolutionary diversification to the rapid evolution of sophisticated visual systems capable of high-bandwidth electromagnetic information processing (Parker, 2003, 2011), suggests that electromagnetic information processing capabilities have been major drivers of evolutionary innovation rather than merely incidental consequences of other selective pressures. Additional support comes from developmental biology research demonstrating that bioelectric patterns can override genetic programs to direct morphogenesis, regeneration, and even induce ectopic organ formation (Levin et al., 2017; Pai et al., 2012; Mathews & Levin, 2018), indicating that electromagnetic field-based information systems have deep evolutionary roots extending beyond neural tissues.
Here we synthesize evidence from multiple disciplines including evolutionary biology, quantum physics, neuroscience, and developmental biology to propose that biological information processing represents progressive optimization of electromagnetic field dynamics across evolutionary time scales. We examine specific evolutionary transitions including the development of chemiosmotic gradient utilization, the emergence of sophisticated photoreception systems during the Cambrian radiation, and the evolution of neural electromagnetic field integration as examples of electromagnetic field optimization under natural selection. We further discuss implications for understanding consciousness as an emergent property of electromagnetic field integration and potential applications in developing electromagnetic-based therapeutic approaches for neurological disorders.
Table 1.
Evolutionary Timeline of Electromagnetic Information Processing.
Table 1.
Evolutionary Timeline of Electromagnetic Information Processing.
| Time Period |
Innovation |
Electromagnetic Mechanism |
Information Advantage |
Organisms |
Speed/Efficiency Gain |
| ~3.5 BYA |
DNA Electromagnetic Stability |
Duplex electromagnetic base pairing |
Enhanced electromagnetic information storage stability |
Early cells |
10× electromagnetic stability |
| ~3.5 BYA |
ATP Synthase Electromagnetic Revolution |
Optimized electromagnetic proton gradient utilization |
3-4× electromagnetic energy extraction efficiency |
Early prokaryotes |
300-400% electromagnetic efficiency |
| ~3.0 BYA |
Proton Network Quantum Jazz |
Coherent electromagnetic water networks and electromagnetic proton cascades |
Near-light-speed electromagnetic information transmission |
Advanced prokaryotes |
~300,000× electromagnetic speed |
| ~2.5 BYA |
Electromagnetic Photosynthesis |
Quantum electromagnetic energy conversion |
Efficient electromagnetic solar energy capture |
Cyanobacteria |
95% electromagnetic conversion efficiency |
| ~1.5 BYA |
Electromagnetic Cellular Integration |
Electromagnetic organelle coordination |
Multi-system electromagnetic integration |
Early eukaryotes |
100× electromagnetic coordination |
| ~550 MYA |
Cephalization Electromagnetic Integration |
Concentrated electromagnetic neural processing |
Spatial electromagnetic integration and electromagnetic specialization |
Planarian flatworms |
10-100× electromagnetic coordination |
| ~540 MYA |
Cambrian Vision Electromagnetic Revolution |
Sophisticated electromagnetic photoreception arrays |
Massive electromagnetic spatial information bandwidth |
Trilobites, early vertebrates |
1,000× electromagnetic information bandwidth |
| ~500 MYA |
Voltage-Gated Electromagnetic Channels |
Action potential electromagnetic signaling |
Rapid long-distance electromagnetic transmission |
Early nervous systems |
1,000× faster than electromagnetic chemical diffusion |
| ~400 MYA |
Electromagnetic Myelination |
Electromagnetic insulation and electromagnetic saltatory conduction |
High-speed electromagnetic neural transmission |
Early vertebrates |
100× electromagnetic transmission speed |
| ~200 MYA |
Electromagnetic Field Integration |
Ephaptic electromagnetic coupling and electromagnetic field-based binding |
Brain-wide instantaneous electromagnetic coordination |
Complex vertebrate brains |
5,000× faster than electromagnetic synaptic transmission |
| ~50 MYA |
Electromagnetic Consciousness Integration |
Coherent electromagnetic field dynamics |
Unified electromagnetic conscious experience |
Advanced mammals |
125 billion× information density advantage |
Table 2.
Information Processing Capabilities by Evolutionary Stage.
Table 2.
Information Processing Capabilities by Evolutionary Stage.
| Stage |
Integration Scope |
Processing Speed |
Information Capacity |
Binding Mechanism |
Evolutionary Bottleneck Solved |
| Chemical Gradients |
Local cellular |
Seconds-minutes |
Limited molecular recognition |
Concentration gradients |
Basic environmental sensing |
| DNA Systems |
Intracellular genetic |
Hours-generations |
Vast genetic storage |
Base-pair complementarity |
Information stability crisis |
| Proton Circuits |
Cellular energetics |
Milliseconds |
Moderate electrochemical |
Chemiosmotic coupling |
Energy-information coupling |
| Water Networks |
Tissue-level coherence |
Nanoseconds |
High quantum information |
Coherent oscillations |
Local field coordination |
| Neurotransmission |
Intercellular signaling |
Milliseconds |
Moderate chemical specificity |
Receptor binding |
Multicellular coordination |
| Cephalization |
Regional neural networks |
Milliseconds |
Enhanced spatial integration |
Neural convergence |
Distributed processing limits |
| Electrical Signaling |
Organism-wide networks |
Milliseconds |
High temporal precision |
Action potential propagation |
Distance-speed constraints |
| EM Field Integration |
Brain-wide consciousness |
Microseconds |
Massive parallel processing |
Electromagnetic field binding |
Consciousness-speed integration |
Table 3.
Quantitative Evolutionary Advantages of Major Transitions.
Table 3.
Quantitative Evolutionary Advantages of Major Transitions.
| Pathway Comparison |
Speed Advantage |
Parallel Processing |
Energy Efficiency |
Information Density |
Integration Scope |
| Chemical → ATP Synthase |
1× |
1× |
3-4× |
1× |
Cellular |
| Chemical → Water Networks |
~10,000,000× |
1,000× |
100× |
1,000× |
Tissue |
| Chemical → Electrical |
1,000× |
10× |
10× |
100× |
Organism |
| Electrical → EM Fields |
5,000× |
10,000× |
1,000× |
10,000× |
Brain-wide |
| Synaptic → Field Integration |
40,000× (latency) |
10,000× |
1,000× |
10,000× |
Consciousness-level |
2. Electromagnetic Field Foundations of Biological Information Processing
All biological processes operate within the framework of fundamental physics, where the four known forces—electromagnetic, strong nuclear, weak nuclear, and gravitational—govern interactions at different scales (Carroll, 2010; Weinberg, 1995). In biological systems, strong and weak nuclear forces operate at sub-atomic scales largely irrelevant to molecular-level information processing, while gravitational effects, though providing essential large-scale environmental organization, contribute negligibly to the molecular-scale electromagnetic interactions that dominate biological information transfer (Atkins & de Paula, 2014; Berg, 1993). Consequently, virtually all biological information processing represents electromagnetic field phenomena operating across different spatial and temporal scales, from molecular recognition events involving electromagnetic interactions between electron orbitals and atomic nuclei, to protein conformational changes resulting from electromagnetic field-driven alterations in bonding patterns, to ion channel function depending on electromagnetic field-gated conformational transitions (Hille, 2001; Dill & MacCallum, 2012).
Chemical bond formation and dissolution, the fundamental currency of molecular information processing, proceed through electromagnetic interactions between charged particles, with covalent bonds forming through electromagnetic sharing of electron pairs, ionic bonds through electromagnetic attraction between oppositely charged atoms, and hydrogen bonds through electromagnetic dipole interactions (Atkins & de Paula, 2014; Pauling, 1960). Protein folding, essential for virtually all biological information processing, results from the minimization of electromagnetic free energy through optimization of electromagnetic interactions between amino acid side chains, backbone atoms, and surrounding water molecules (Dill & MacCallum, 2012; Anfinsen, 1973). Enzyme catalysis operates through electromagnetic field effects that stabilize transition states and orient reactant molecules through precisely tuned electromagnetic environments within active sites (Warshel et al., 2006; Kamerlin & Warshel, 2010). Even large-scale biological processes such as morphogenesis, regenerative responses, and developmental patterning operate through bioelectric field coordination of cellular behaviors, where electromagnetic fields created by ion flux patterns provide positional information and coordinate multicellular responses (Levin et al., 2017; McCaig et al., 2005; Zhao, 2009).
Information theory, originally developed for analyzing communication systems (Shannon, 1948), provides quantitative frameworks for understanding biological information processing efficiency and capacity (Adami, 2004; Walker et al., 2017; Still et al., 2012). Electromagnetic field-based information transmission offers several fundamental theoretical advantages over purely chemical mechanisms that may explain evolutionary trends toward electromagnetic field-based biological computation. Electromagnetic field propagation approaches light speed (3×10⁸ m/s) compared to molecular diffusion proceeding at approximately 10⁻⁶ m²/s, yielding theoretical speed advantages of up to 14 orders of magnitude for field-based versus chemical information transmission (Berg, 1993; Ruffini et al., 2020). Electromagnetic fields naturally support multiple frequency channels simultaneously through wave superposition, enabling parallel information transmission and processing, while chemical signaling systems are typically limited to sequential molecular recognition events with inherent temporal bottlenecks (Ho, 1998; Lisman & Jensen, 2013; Buzsaki, 2006). The linear superposition property of electromagnetic fields enables natural integration of multiple information streams through field summation, while chemical systems require complex enzymatic cascades for signal integration with associated metabolic costs and temporal delays (Mitchell, 1961; McFadden, 2020; Katz & Miledi, 1970). Furthermore, electromagnetic field coherence enables redundant encoding and quantum error correction through field interference patterns, potentially providing more robust information storage and transmission compared to chemical systems that rely on molecular proofreading mechanisms with inherent error rates (Lambert et al., 2013; Ho, 2008; Zurek, 2003).
3. Evolutionary Optimization of Electromagnetic Information Processing
The evolutionary transition from RNA-based to DNA-based information storage approximately 3.5 billion years ago represents an early example of natural selection optimizing electromagnetic field properties for enhanced information processing and storage (Lazcano et al., 1988; Joyce, 2002; Freeland & Hurst, 1998). DNA's double-helix structure provides enhanced electromagnetic stability through multiple mechanisms that collectively improve information fidelity and longevity compared to single-stranded RNA precursors. Watson-Crick base pairing creates complementary electromagnetic charge distributions that stabilize information storage through hydrogen bonding networks, which are themselves electromagnetic interactions between electronegative atoms and hydrogen atoms carrying partial positive charges (Watson & Crick, 1953; Franklin & Gosling, 1953; Saenger, 1984). The double-strand structure provides redundant information encoding that enables electromagnetic error detection and correction, as mismatched bases create electromagnetic field distortions detectable by repair enzymes that have evolved exquisite sensitivity to electromagnetic anomalies in DNA structure (Friedberg et al., 2006; Lindahl, 1993; Sancar, 1996). Additionally, the double-helix configuration provides electromagnetic shielding effects that substantially reduce susceptibility to UV radiation damage compared to single-strand RNA, with the electromagnetic properties of stacked base pairs creating protective electromagnetic field configurations (Bernstein et al., 1999; Sagan, 1973; Cadet et al., 2005).
Quantitative analyses comparing DNA and RNA stability under physiological conditions demonstrate that DNA provides approximately 10-fold greater electromagnetic stability than RNA, with significantly reduced rates of spontaneous hydrolysis, oxidative damage, and UV-induced lesion formation (Joyce, 2002; Poole et al., 1998; Lindahl, 1993). This substantial improvement in electromagnetic information storage capacity would have provided significant selective advantages for early organisms, driving the evolutionary transition from RNA-based to DNA-based information systems despite the additional metabolic costs associated with DNA synthesis and maintenance (Lane, 2015; Koonin & Martin, 2005). The rapid fixation of DNA-based information storage across all domains of life suggests that electromagnetic field optimization was a major evolutionary driving force even in early biological systems.
The evolution of ATP synthase represents another critical transition in biological electromagnetic energy and information processing, demonstrating how natural selection can optimize complex molecular machines for enhanced electromagnetic efficiency (Mitchell, 1961; Boyer, 1997; Lane, 2015). This remarkable molecular motor exploits electromagnetic proton gradients across biological membranes to drive ATP synthesis, achieving 3-4 fold greater energy extraction efficiency compared to simpler chemiosmotic systems that preceded it (Walker, 2013; Boyer, 1993; Yoshida et al., 2001). The mechanism involves sophisticated electromagnetic field effects, where proton concentration differences create electromagnetic potential gradients that drive conformational changes in the ATP synthase complex through precisely coordinated electromagnetic interactions between charged amino acid residues, metal ions, and substrate molecules (Abrahams et al., 1994; Stock et al., 1999). The evolutionary optimization of this system demonstrates how complex electromagnetic field dynamics can be harnessed to dramatically improve biological energy processing efficiency, with implications extending beyond metabolism to information processing systems that depend on available energy for operation.
Mae-Wan Ho's groundbreaking theoretical framework proposed that biological systems exhibit "quantum jazz," characterized by coherent electromagnetic oscillations that enable rapid information transfer through structured water networks exhibiting liquid crystalline properties throughout living organisms (Ho, 1998, 2008; Ho & Knight, 1998). This framework suggests that proton concentration gradients, beyond their role in energy production, function as sophisticated electromagnetic information carriers capable of encoding and transmitting complex signals across cellular and tissue boundaries through coordinated electromagnetic field changes and electrochemical cascade reactions (Ho, 1993, 1995). Experimental evidence supporting this hypothesis includes observations of coherent biophoton emission from living tissues, long-range electromagnetic coherence in cellular systems, and the sensitivity of biological processes to extremely weak electromagnetic fields that would be inconsistent with purely chemical signaling mechanisms (Popp et al., 1992; Fröhlich, 1988; Del Giudice et al., 2005). Ho's work further proposed that structured water in biological systems exhibits electromagnetic properties that support coherent oscillations and near-instantaneous information propagation across biological tissues through electromagnetic field effects rather than molecular diffusion (Ho, 1998; Ho & Knight, 1998). This water-based electromagnetic information transfer system would operate in parallel with conventional biochemical signaling pathways, potentially explaining the remarkable speed and coordination observed in complex biological responses that appear to exceed the capabilities of purely chemical communication systems.
Subsequent research has provided additional support for electromagnetic information processing in biological proton networks, including demonstrations of quantum coherence in biological proton transport systems, long-range electromagnetic correlations in living tissues, and the ability of extremely weak electromagnetic fields to influence biological processes in ways that suggest electromagnetic field-based rather than chemical signaling mechanisms (Lambert et al., 2013; Penrose, 1989; Hameroff & Penrose, 2014). The evolutionary implications of such systems are profound, as organisms capable of electromagnetic field-based information processing would possess significant advantages in environmental responsiveness, internal coordination, and energy efficiency compared to organisms limited to purely chemical signaling (Ho, 2008; McFadden, 2020). The apparent conservation of these electromagnetic properties across diverse biological systems, from bacterial communities to neural networks, suggests that electromagnetic field optimization has been a persistent evolutionary theme throughout the history of life.
4. The Cambrian Electromagnetic Vision Revolution
Andrew Parker's "Light Switch" hypothesis proposes that the rapid evolution of sophisticated visual systems during the Cambrian explosion (approximately 540-520 million years ago) represents a pivotal transition in biological electromagnetic information processing that fundamentally transformed Earth's evolutionary dynamics (Parker, 2003, 2011; Marshall & Ahlberg, 2014). This hypothesis suggests that the emergence of advanced eyes capable of high-bandwidth electromagnetic information processing was not merely a consequence of existing evolutionary diversification but rather a primary driving force that created entirely new selective pressures and ecological opportunities through dramatically enhanced information gathering capabilities. The fossil record provides compelling evidence for this hypothesis, with sophisticated compound eyes appearing in trilobites within a relatively brief evolutionary timeframe, demonstrating remarkable optical engineering that achieved near-optimal electromagnetic information processing through precisely aligned calcite lenses and photoreceptor arrays (Clarkson & Levi-Setti, 1975; Gal et al., 2000; Schoenemann & Clarkson, 2013).
Trilobite compound eyes represent extraordinary examples of evolutionary electromagnetic engineering, with individual calcite lenses arranged in hexagonal arrays that minimize optical aberrations and maximize electromagnetic information gathering efficiency across wide visual fields (Clarkson & Levi-Setti, 1975; Lee et al., 2007). The optical properties of these systems approach theoretical limits for electromagnetic information processing, with calcite's unique birefringent properties enabling sophisticated electromagnetic focusing that eliminates spherical aberration—a feat not achieved in human-designed optical systems until the 17th century work of Huygens and Descartes (Gal et al., 2000). Each compound eye contained hundreds of individual photoreceptive units capable of parallel electromagnetic information processing, creating visual systems with extraordinary sensitivity to movement, spatial patterns, and electromagnetic environmental changes (Schoenemann & Clarkson, 2013; Fortey & Owens, 1999). The rapid evolutionary appearance of such sophisticated electromagnetic information processing systems suggests intense selective pressure for enhanced visual capabilities, consistent with Parker's hypothesis that electromagnetic vision fundamentally altered evolutionary dynamics.
The electromagnetic information processing revolution enabled by sophisticated vision created cascading evolutionary effects that may explain the remarkable diversification characteristic of the Cambrian explosion (Erwin & Valentine, 2013; Marshall, 2006). Visual predators gained unprecedented ability to detect and track prey across three-dimensional environments through electromagnetic pattern recognition, creating intense selective pressure for defensive innovations including armor, camouflage, and rapid escape responses (Bengtson, 2002; Conway Morris, 1998). The electromagnetic information gathered through vision enabled complex behavioral innovations including coordinated group behaviors, sophisticated predator-prey interactions, and exploitation of previously inaccessible ecological niches requiring precise electromagnetic navigation and environmental assessment (Parker, 2003; Paterson et al., 2011). Additionally, visual communication systems emerged that exploited electromagnetic signaling through color patterns, bioluminescence, and morphological displays, creating new dimensions of biological information exchange that required sophisticated neural electromagnetic information processing for both signal production and interpretation (Nilsson, 2013; Arendt, 2003).
The evolution of camera eyes in early vertebrates represents an alternative solution to electromagnetic information processing optimization, achieving high-resolution electromagnetic imaging through different optical principles than compound eye systems (Nilsson & Pelger, 1994; Fernald, 2006). Camera eyes focus electromagnetic radiation through single lens systems to create high-resolution images on photoreceptive retinal surfaces, enabling detailed electromagnetic pattern recognition and precise electromagnetic environmental monitoring (Land & Nilsson, 2012). The evolutionary optimization of camera eye systems demonstrates remarkable convergent evolution, with similar electromagnetic focusing principles evolving independently in vertebrates, cephalopods, and other lineages, suggesting strong selective pressure for high-performance electromagnetic information processing systems (Nilsson, 2004; Packard, 1972). The neural processing requirements for interpreting electromagnetic information from camera eyes necessitated the evolution of sophisticated visual cortical systems capable of real-time electromagnetic pattern analysis, edge detection, motion processing, and three-dimensional electromagnetic scene reconstruction (Hubel & Wiesel, 1962; Livingstone & Hubel, 1988).
The electromagnetic information processing capabilities enabled by Cambrian visual systems created evolutionary pressure for increasingly sophisticated neural integration systems capable of processing multiple electromagnetic information streams simultaneously while coordinating complex behavioral responses (Arendt, 2003; Strausfeld & Hirth, 2013). This pressure likely contributed to the evolution of centralized nervous systems with enhanced electromagnetic information integration capabilities, laying the foundation for later innovations in electromagnetic field-based neural processing that would eventually give rise to complex cognition and consciousness (Holland, 2003; Northcutt, 2012). The Cambrian electromagnetic vision revolution thus represents a critical transition point where biological electromagnetic information processing transcended simple molecular recognition to achieve sophisticated electromagnetic field analysis and integration, establishing evolutionary trajectories that would culminate in the electromagnetic field-based neural processing characteristic of advanced cognition.
5. Cephalization and Electromagnetic Neural Integration
The evolutionary trend toward cephalization, involving the concentration of sensory organs and neural processing centers in anterior body regions, represents a fundamental reorganization of biological electromagnetic information processing that emerged approximately 550 million years ago in early bilaterian organisms (Arendt, 2021; Holland, 2003; Northcutt, 2012). This transition, first clearly evident in planarian flatworms and subsequently elaborated across diverse animal lineages, reflects optimization of electromagnetic information processing through spatial concentration of neural elements and sensory systems (Cebrià, 2008; Reddien & Alvarado, 2004). Cephalization enabled several critical advances in electromagnetic information processing including compartmentalization and specialization of neural functions allowing different brain regions to optimize for specific types of electromagnetic signal processing, integration of multiple sensory electromagnetic inputs enabling rapid multimodal information analysis, and concentration of neural density facilitating electromagnetic field interactions between closely spaced neural elements (Strausfeld & Hirth, 2013; Tomer et al., 2010).
The evolutionary advantages of cephalized electromagnetic information processing become apparent when considering the physical constraints of biological signaling systems and the benefits of spatial organization for electromagnetic field generation and coordination (Young et al., 2024; Arendt, 2021). Concentrated neural populations create optimal conditions for generating coherent electromagnetic fields through synchronized electrical activity across densely packed neural networks, enabling electromagnetic field-based information integration that transcends the limitations of individual synaptic connections (McFadden, 2020; Hunt & Jones, 2023). The anterior positioning of sensory organs in cephalized organisms optimizes electromagnetic information gathering by concentrating detection systems at the leading edge of locomotion, enabling rapid assessment of environmental electromagnetic cues and coordinated behavioral responses through integrated neural processing (Holland, 2003; Arendt, 2003). Additionally, the spatial segregation of different neural functions within cephalized brains enables parallel electromagnetic information processing streams that can operate simultaneously while maintaining the capability for higher-order electromagnetic integration when required for complex behavioral responses.
Phylogenetic analyses reveal that cephalization evolved independently multiple times across animal lineages, suggesting strong selective pressure for concentrated electromagnetic information processing systems (Strausfeld & Hirth, 2013; Moroz, 2009; Northcutt, 2012). Comparative studies of cephalized and non-cephalized nervous systems demonstrate clear performance advantages for cephalized organizations in tasks requiring rapid electromagnetic information integration, complex electromagnetic pattern recognition, and coordinated electromagnetic motor responses (Farris, 2016; Chittka & Niven, 2009). The evolutionary persistence and elaboration of cephalized neural organizations across diverse environmental conditions and ecological niches indicates that electromagnetic information processing optimization through spatial neural concentration provides robust adaptive advantages that transcend specific environmental pressures.
The concentration of neural elements during cephalization creates physical conditions conducive to electromagnetic field generation and coordination through synchronized neural activity across spatially proximate cells (Buzsaki, 2006; Wang, 2010). When large numbers of neurons fire synchronously within confined brain regions, their individual electromagnetic fields summate to create macroscopic electromagnetic field effects that can influence neural activity across the entire brain region through electromagnetic field interactions rather than requiring specific synaptic connections (McFadden, 2020; Anastassiou et al., 2011). This electromagnetic field-based neural coordination provides a physical mechanism for rapid information integration across distributed neural networks, enabling real-time electromagnetic binding of diverse information streams into coherent neural representations (Engel & Singer, 2001; Ward, 2003).
6. Ephaptic Coupling and Electromagnetic Field-Based Neural Processing
The discovery and characterization of ephaptic coupling represents a paradigm shift in understanding neural electromagnetic information processing, revealing that direct electromagnetic field interactions between neurons can transmit information at speeds approaching 47 km/s—nearly 500 times faster than the fastest myelinated axons and 5,000 times faster than unmyelinated neural transmission (Anastassiou et al., 2011; Pinotsis & Miller, 2023; Chiang et al., 2019). Ephaptic coupling, from the Greek "ephapsis" meaning "to touch upon," describes electromagnetic field interactions between adjacent neural elements that occur without direct synaptic connections, enabling rapid electromagnetic information transfer through field effects rather than chemical transmission (Arvanitaki, 1942; Katz & Schmitt, 1940). This mechanism operates through several distinct electromagnetic processes including direct electromagnetic field coupling between neural processes enabling instantaneous information transfer, electromagnetic field-based synchronization of neural oscillations creating coherent electromagnetic states across brain regions, and electromagnetic field summation effects that enable non-linear electromagnetic information integration across distributed neural networks.
Recent experimental advances have provided compelling evidence for direct electromagnetic field influence on neural activity, challenging traditional assumptions about information flow direction in neural systems. Anastassiou and colleagues demonstrated that weak electromagnetic fields (≥0.74 mV/mm) can reliably entrain neural spike timing, establishing causal field-to-neuron communication pathways that operate below the threshold of conventional synaptic transmission (Anastassiou et al., 2011). This finding suggests that electromagnetic fields can directly modulate neural computation rather than serving merely as passive byproducts of neural activity.
Table 4.
Electromagnetic Field vs Neural Spike Processing Comparison.
Table 4.
Electromagnetic Field vs Neural Spike Processing Comparison.
| Processing Parameter |
Electromagnetic Fields |
Neural Spikes |
Advantage Factor |
References |
| Propagation Speed |
47-57 km/s (brain tissue) |
0.1-120 m/s |
500-5,000× |
Ruffini et al., 2020; Anastassiou et al., 2011 |
| Information Density |
Volumetric (3D) processing |
Linear pathway processing |
125 billion× (theoretical) |
Hunt & Jones, 2023; Pinotsis & Miller, 2023 |
| Parallel Processing |
All-to-all field coupling |
Limited synaptic connections |
10,000× simultaneous operations |
Hameroff & Penrose, 2014 |
| Response Latency |
<10⁻⁸ seconds (ephaptic) |
1-20 ms (synaptic) |
40,000× faster response |
Chiang et al., 2019 |
| Communication Range |
mm-cm scale coherence |
μm synaptic gaps |
1,000× greater range |
Francis et al., 2003 |
| Energy Efficiency |
~10⁻¹⁸ J/operation |
~10⁻¹⁵ J/operation |
1,000× more efficient |
Hunt & Schooler, 2019 |
| Bandwidth Capacity |
THz theoretical range |
kHz firing frequencies |
10,000× information flow |
Bandyopadhyay, 2019 |
| Entrainment Threshold |
0.74 mV/mm |
Several mV depolarization |
5,000× sensitivity |
Anastassiou et al., 2011 |
| Synchronization Options |
Sub-Hz to THz frequencies |
Limited by refractory periods |
1,000,000× coordination |
Freeman, 2004 |
| Integration Capability |
Instantaneous field effects |
Sequential synaptic delays |
Continuous vs discrete |
McFadden, 2020 |
Electrocorticography (ECoG) studies have revealed patterns that strongly support the strong electromagnetic field hypothesis, where electromagnetic fields constitute the primary computational substrate while neural spikes provide energetic drive. The critical observation that broadband ECoG power correlates more strongly with local neural population activity than oscillatory patterns fundamentally challenges spike-centric computational models (Brake et al., 2024). If neural spikes were the primary computational mechanism, ECoG recordings should reveal oscillatory patterns reflecting spike timing codes and narrow-band frequency signatures corresponding to specific computational processes. Instead, ECoG consistently demonstrates broadband power increases suggesting continuous electromagnetic field computation with spatial specificity that exceeds what spike timing alone could explain (Finley et al., 2024).
The significance of aperiodic neural activity, previously dismissed as "neural noise," has emerged as a crucial component of electromagnetic field-based neural computation. Brake and colleagues demonstrated through detailed biophysical modeling that aperiodic EEG activity reflects functionally meaningful brain states arising from complex network dynamics and arrhythmic neural activity (Brake et al., 2024). Most remarkably, Finley and colleagues found that aperiodic exponents measured during resting states could predict cognitive decline over 10-year periods, with the aperiodic component reflecting the balance between excitatory and inhibitory neural activity (Finley et al., 2024). This predictive power suggests that electromagnetic field patterns capture fundamental aspects of brain health and cognitive capacity that conventional spike-based measures fail to detect.
The evolutionary implications of ephaptic coupling are profound, as electromagnetic field-based neural processing provides significant advantages in speed, integration capability, and energy efficiency compared to purely synaptic mechanisms (Hunt & Jones, 2023; McFadden, 2020). Neural systems capable of electromagnetic field-based information processing can achieve real-time integration of information across entire brain regions, enabling rapid decision-making and coordinated responses that would be impossible with synaptic transmission alone due to temporal delays inherent in chemical signaling (Freeman, 2004; John, 2002). The ability to process multiple information streams simultaneously through electromagnetic field superposition enables parallel processing capabilities that dramatically expand computational capacity compared to sequential synaptic processing (Lisman & Jensen, 2013; Buzsaki & Wang, 2012). Additionally, electromagnetic field-based processing requires significantly less metabolic energy than maintaining extensive synaptic networks, as electromagnetic field effects operate through field interactions rather than requiring continuous synthesis and recycling of neurotransmitter molecules (Laughlin et al., 1998; Lennie, 2003).
Recent research has demonstrated that ephaptic coupling plays critical roles in memory formation, with electromagnetic field interactions enabling rapid consolidation of information across neural networks without requiring slow synaptic plasticity mechanisms (Pinotsis & Miller, 2023; Tsodyks, 1999). Electromagnetic field-based memory formation can occur within milliseconds compared to the minutes or hours required for synaptic plasticity, suggesting that ephaptic mechanisms may be responsible for the rapid learning and memory formation observed in complex behavioral tasks (Martin et al., 2000; Kandel, 2001). Furthermore, electromagnetic field-based neural processing provides a natural mechanism for binding distributed neural representations into unified perceptual and cognitive experiences, addressing the long-standing binding problem in neuroscience through electromagnetic field integration rather than requiring complex synaptic connectivity patterns (Engel & Singer, 2001; Revonsuo & Newman, 1999).
7. Consciousness as Electromagnetic Field Integration
The emergence of consciousness represents perhaps the most sophisticated achievement in biological electromagnetic information processing, involving the integration of multiple electromagnetic pathways operating simultaneously across diverse spatial and temporal scales (McFadden, 2020; Hunt & Jones, 2023; Hameroff & Penrose, 2014). Contemporary theories of consciousness increasingly recognize that conscious experience requires information integration capabilities that exceed the temporal and spatial constraints of purely synaptic mechanisms, suggesting fundamental roles for electromagnetic field-based neural processing in generating unified conscious states (Tononi, 2004; Dehaene, 2014; Edelman & Tononi, 2000). The binding problem, which asks how the brain integrates diverse sensory and cognitive information into unified conscious experience, finds natural solutions in electromagnetic field theories where information can be integrated instantaneously across entire brain regions through field effects rather than requiring sequential synaptic transmission with inherent temporal delays (Revonsuo & Newman, 1999; Treisman, 1996).
Electromagnetic field theories of consciousness propose that conscious experience emerges from coherent electromagnetic field dynamics in neural tissue, where synchronized neural activity creates macroscopic electromagnetic field states that integrate information across distributed brain regions (McFadden, 2020; Pockett, 2000; John, 2002). These theories suggest that consciousness represents a phase transition in neural electromagnetic organization, where locally processed information becomes globally accessible through electromagnetic field integration mechanisms that transcend anatomical connectivity constraints (Edelman & Tononi, 2000; Baars, 1988). The temporal characteristics of conscious experience, including the unity of conscious moments and the continuity of conscious states across time, align with electromagnetic field dynamics that can maintain coherent states across biologically relevant timescales while enabling rapid transitions between different conscious contents (Freeman, 2004; VanRullen & Koch, 2003).
Experimental evidence supporting electromagnetic field theories of consciousness includes observations that transcranial electromagnetic stimulation can reliably modulate conscious experience, suggesting direct electromagnetic field influences on consciousness (Rossi et al., 2009; Hallett, 2007). Electromagnetic field recordings from neural tissue demonstrate coherent field oscillations that correlate with conscious states and cognitive processes, with disruption of these electromagnetic patterns leading to alterations in conscious experience (Buzsaki, 2006; Llinas et al., 1998). Additionally, electromagnetic field-based theories provide parsimonious explanations for the remarkable speed of conscious processing, the unity of conscious experience across diverse sensory modalities, and the global accessibility of conscious information throughout the nervous system (Freeman, 2004; John, 2002).
7.2. Electromagnetic Field Solution to the Binding Problem
The binding problem—how the brain integrates distributed information processing into unified conscious experience—finds natural resolution through electromagnetic field theory rather than requiring complex synaptic connectivity patterns (Revonsuo & Newman, 1999; Treisman, 1996). Traditional computational approaches have struggled with this problem because they attempt to artificially connect discrete neural modules through precise temporal coordination, yet empirical evidence consistently fails to support the predicted synchronization mechanisms required for spike-based binding (Engel & Singer, 2001; Ward, 2003).
Electromagnetic field-based binding operates through fundamentally different mechanisms that exploit the wave properties of electromagnetic fields. Through superposition, multiple field patterns can coexist and interact simultaneously across brain regions without requiring discrete anatomical connections (McFadden, 2020; Hunt & Jones, 2023). Cross-frequency coupling enables different oscillatory frequencies to nest within each other, creating hierarchical information integration across multiple spatial and temporal scales (Canolty & Knight, 2010; Lisman & Jensen, 2013). Phase relationships and resonance patterns allow distant brain regions to achieve rapid coordination that consciousness requires while maintaining the flexible, context-sensitive integration that conscious experience exhibits (Freeman, 2004; Buzsaki, 2006).
The elegance of electromagnetic field-based binding emerges from its ability to solve both spatial and temporal aspects of the binding problem through a single mechanism. Spatial binding occurs through electromagnetic field patterns that naturally span brain regions, creating unified representational states without requiring convergence zones where different processing streams must artificially meet (Freeman, 1991; John, 2002). Temporal binding occurs through electromagnetic field dynamics that can maintain coherent states across biologically relevant timescales while enabling rapid transitions between different conscious contents (VanRullen & Koch, 2003; Ward, 2003). Unlike spike-based approaches that must construct unity from discrete elements, electromagnetic field binding preserves the natural unity of conscious experience while explaining how distributed neural processes contribute to unified conscious states.
8. Quantum Biology and Electromagnetic Coherence
The emerging field of quantum biology has revealed that quantum electromagnetic effects play functional roles in biological information processing, challenging previous assumptions that quantum coherence cannot persist in warm, wet biological environments (Lambert et al., 2013; McFadden & Al-Khalili, 2018; Marais et al., 2018). Discoveries of quantum coherence in photosynthetic light harvesting complexes, where quantum superposition enables near-perfect energy transfer efficiency through optimization of electromagnetic energy transport pathways, demonstrate that biological systems have evolved mechanisms to harness quantum electromagnetic effects for enhanced performance (Engel et al., 2007; Collini et al., 2010; Panitchayangkoon et al., 2010). Similar quantum electromagnetic effects have been identified in avian magnetoreception, where cryptochrome proteins in retinal neurons exhibit quantum entanglement that enables detection of extremely weak electromagnetic field variations associated with Earth's magnetic field (Ritz et al., 2000; Mouritsen & Ritz, 2005; Hore & Mouritsen, 2016).
The persistence of quantum electromagnetic coherence in biological systems requires sophisticated electromagnetic protection mechanisms that shield quantum states from environmental decoherence while maintaining functional interactions with cellular environments (Tegmark, 2000; Hagan et al., 2002). Biological quantum coherence appears to exploit specific electromagnetic properties of protein structures and water networks that create protected electromagnetic environments where quantum effects can persist for biologically relevant timescales (Lambert et al., 2013; Chin et al., 2013). These quantum electromagnetic protection mechanisms represent remarkable evolutionary innovations that enable biological systems to exploit quantum mechanical properties for enhanced information processing and energy transfer efficiency.
Hameroff and Penrose have proposed that quantum electromagnetic effects in neural microtubules may contribute to consciousness through orchestrated objective reduction, where quantum superposition states in microtubule protein networks undergo objective reduction events that generate conscious moments (Hameroff & Penrose, 2014; Penrose, 1989). This theory suggests that consciousness involves quantum electromagnetic information processing that transcends classical neural computation, with quantum superposition enabling simultaneous exploration of multiple information processing pathways that are resolved into specific conscious contents through quantum electromagnetic reduction events (Hameroff, 2006; Tegmark, 2000). While controversial, this hypothesis has gained support from demonstrations of quantum electromagnetic effects in biological systems and discoveries of sophisticated electromagnetic structures in neural microtubules that could potentially support quantum coherence (Bandyopadhyay, 2019; Craddock et al., 2015).
The evolutionary implications of quantum biology suggest that natural selection has repeatedly discovered and optimized quantum electromagnetic mechanisms to enhance biological performance beyond classical limits (McFadden & Al-Khalili, 2018; Davies, 2004). The conservation of quantum electromagnetic effects across diverse biological systems from photosynthetic bacteria to neural networks indicates that quantum biology represents a fundamental aspect of biological organization rather than isolated evolutionary innovations (Lambert et al., 2013; Marais et al., 2018). This perspective suggests that understanding biological information processing requires integration of quantum electromagnetic effects with classical electromagnetic mechanisms, revealing biological systems as sophisticated quantum electromagnetic processors that have been optimized by evolution over billions of years.
9. Applications and Future Directions
The electromagnetic field perspective on biological information processing has significant implications for understanding neurological disorders and developing therapeutic interventions based on electromagnetic field modulation (Hunt, 2024; Bestmann & Walsh, 2017). Many neurological and psychiatric conditions may involve disruptions in electromagnetic field coordination rather than purely synaptic dysfunction, suggesting that electromagnetic field-based therapies could provide more effective treatments than conventional pharmacological approaches targeting individual neurotransmitter systems (Rossi et al., 2009; George et al., 2010). Transcranial electromagnetic stimulation techniques, including transcranial magnetic stimulation and transcranial direct current stimulation, have demonstrated therapeutic efficacy for depression, chronic pain, and cognitive disorders through direct modulation of neural electromagnetic fields (Hallett, 2007; Nitsche et al., 2008).
The development of bioengineering applications based on electromagnetic biological principles represents another promising research direction, with potential applications including bio-inspired computing systems that integrate multiple electromagnetic information processing pathways, electromagnetic communication technologies based on biological field coordination mechanisms, and electromagnetic enhancement systems that optimize human cognitive performance through field modulation (Hunt & Jones, 2023; McFadden, 2020). Understanding biological electromagnetic field processing could inform the development of artificial intelligence systems that achieve consciousness-like properties through electromagnetic field integration rather than purely digital computation (Hameroff, 2006; Tegmark, 2000).
Future research directions should focus on detailed electromagnetic field mapping of biological systems across multiple spatial and temporal scales, investigation of electromagnetic interactions between different information processing pathways, and development of sophisticated mathematical models of electromagnetic multi-pathway integration (Einevoll et al., 2013; Buzsaki et al., 2012). Advanced experimental techniques including high-resolution electromagnetic field recording, controlled electromagnetic field manipulation, and quantum electromagnetic measurement technologies will be essential for testing electromagnetic theories of biological information processing and consciousness (Anastassiou et al., 2011; Pinotsis & Miller, 2023).
10. Conclusions
The evidence reviewed here supports a unified electromagnetic field perspective on biological information processing in evolution, revealing that apparent mechanistic diversity in biological systems represents different organizational scales of fundamental electromagnetic dynamics rather than entirely distinct physical processes. The evolutionary progression from simple molecular electromagnetic interactions through Ho's quantum jazz and electromagnetic proton networks, Parker's Cambrian electromagnetic vision revolution, to sophisticated electromagnetic field processing in neural systems demonstrates consistent biological optimization of electromagnetic information processing capabilities across evolutionary time.
This electromagnetic framework provides novel insights into longstanding biological questions including the binding problem in neuroscience, the remarkable speed and coordination of biological responses, and the emergence of consciousness as an emergent property of electromagnetic field integration.
The quantitative advantages of electromagnetic field-based information processing—including orders of magnitude improvements in processing speed, parallel processing capabilities, and energy efficiency—suggest that electromagnetic field optimization has been a major evolutionary driving force throughout the history of life. This is not a surprise – EM fields are, after all, the ocean in which all life exists at all times. The conservation of electromagnetic processing mechanisms across diverse biological systems from bacterial communities to human brains indicates that electromagnetic field dynamics represent fundamental organizing principles of biological systems rather than incidental consequences of other evolutionary pressures.
Understanding biological information processing as electromagnetic field dynamics opens new possibilities for therapeutic interventions, bioengineering applications, and artificial intelligence development based on biological electromagnetic principles. This perspective may ultimately provide a unified physical framework for understanding the remarkable sophistication of biological information processing and the emergence of consciousness as the most advanced electromagnetic field integration system evolved on Earth.
References
- Abrahams, J. P.; Leslie, A. G.; Lutter, R.; Walker, J. E. Structure at 2.8 Å resolution of F1-ATPase from bovine heart mitochondria. Nature 1994, 370(6491), 621–628. [Google Scholar] [CrossRef]
- Adami, C. Information theory in molecular biology. Physics of Life Reviews 2004, 1(1), 3–22. [Google Scholar] [CrossRef]
- Anastassiou, C. A.; Perin, R.; Markram, H.; Koch, C. Ephaptic coupling of cortical neurons. Nature Neuroscience 2011, 14(2), 217–223. [Google Scholar] [CrossRef]
- Anfinsen, C. B. Principles that govern the folding of protein chains. Science 1973, 181(4096), 223–230. [Google Scholar] [CrossRef]
- Arendt, D. Evolution of eyes and photoreceptor cell types. International Journal of Developmental Biology 2003, 47(7-8), 563–571. [Google Scholar]
- Arendt, D. Elementary nervous systems. Philosophical Transactions of the Royal Society B 2021, 376(1821), 20200347. [Google Scholar] [CrossRef] [PubMed]
- Arvanitaki, A. Effects evoked in an axon by the activity of a contiguous one. Journal of Neurophysiology 1942, 5(2), 89–108. [Google Scholar] [CrossRef]
- Atkins, P.; de Paula, J. Physical Chemistry: Thermodynamics, Structure, and Change; W. H. Freeman and Company, 2014. [Google Scholar]
- Baars, B. J. A Cognitive Theory of Consciousness; Cambridge University Press, 1988. [Google Scholar]
- Bandyopadhyay, A. Experimental studies on a single microtubule (2013-2019): A brief review. Journal of Physics Communications 2019, 3(3), 035019. [Google Scholar]
- Bengtson, S. Origins and early evolution of predation. Paleontological Society Papers 2002, 8, 289–317. [Google Scholar] [CrossRef]
- Berg, H. C. Random Walks in Biology; Princeton University Press, 1993. [Google Scholar]
- Bernstein, C.; Bernstein, H.; Payne, C. M.; Garewal, H. DNA repair/pro-apoptotic dual-role proteins in five major DNA repair pathways: fail-safe protection against carcinogenesis. Mutation Research 1999, 511(2), 145–178. [Google Scholar] [CrossRef]
- Bestmann, S.; Walsh, V. Transcranial electrical stimulation. Current Biology 2017, 27(23), R1258–R1262. [Google Scholar] [CrossRef]
- Bokil, H.; Laaris, N.; Blinder, K.; Ennis, M.; Keller, A. Ephaptic interactions in the mammalian olfactory system. Journal of Neuroscience 2001, 21(20), RC173. [Google Scholar] [CrossRef] [PubMed]
- Boyer, P. D. The binding change mechanism for ATP synthase—some probabilities and possibilities. Biochimica et Biophysica Acta 1993, 1140(3), 215–250. [Google Scholar] [CrossRef] [PubMed]
- Boyer, P. D. The ATP synthase—a splendid molecular machine. Annual Review of Biochemistry 1997, 66(1), 717–749. [Google Scholar] [CrossRef] [PubMed]
- Buzsaki, G. Rhythms of the Brain; Oxford University Press, 2006. [Google Scholar]
- Buzsaki, G.; Anastassiou, C. A.; Koch, C. The origin of extracellular fields and currents—EEG, ECoG, LFP and spikes. Nature Reviews Neuroscience 2012, 13(6), 407–420. [Google Scholar] [CrossRef]
- Buzsaki, G.; Wang, X. J. Mechanisms of gamma oscillations. Annual Review of Neuroscience 2012, 35, 203–225. [Google Scholar] [CrossRef]
- Cadet, J.; Sage, E.; Douki, T. Ultraviolet radiation-mediated damage to cellular DNA. Mutation Research 2005, 571(1-2), 3–17. [Google Scholar]
- Carroll, S. The Big Picture: On the Origins of Life, Meaning, and the Universe Itself; Dutton, 2010. [Google Scholar]
- Cebrià, F. Organization of the nervous system in the model planarian Schmidtea mediterranea: an immunocytochemical study. Neuroscience Research 2008, 61(4), 375–384. [Google Scholar] [CrossRef]
- Chiang, C. C.; Shivacharan, R. S.; Wei, X.; Gonzalez-Reyes, L. E.; Durand, D. M. Slow periodic activity in the longitudinal hippocampal slice can self-propagate non-synaptically by a mechanism consistent with ephaptic coupling. Journal of Physiology 2019, 597(1), 249–269. [Google Scholar] [CrossRef]
- Chin, A. W.; Datta, A.; Caruso, F.; Huelga, S. F.; Plenio, M. B. Noise-assisted energy transfer in quantum networks and light-harvesting complexes. New Journal of Physics 2013, 12(6), 065002. [Google Scholar] [CrossRef]
- Chittka, L.; Niven, J. Are bigger brains better? Current Biology 2009, 19(21), R995–R1008. [Google Scholar] [CrossRef] [PubMed]
- Clarkson, E.; Levi-Setti, R. Trilobite eyes and the optics of Des Cartes and Huygens. Nature 1975, 254(5502), 663–667. [Google Scholar] [CrossRef] [PubMed]
- Collini, E.; Wong, C. Y.; Wilk, K. E.; Curmi, P. M.; Brumer, P.; Scholes, G. D. Coherently wired light-harvesting in photosynthetic marine algae at ambient temperature. Nature 2010, 463(7281), 644–647. [Google Scholar] [CrossRef] [PubMed]
- Conway Morris, S. The Crucible of Creation: The Burgess Shale and the Rise of Animals; Oxford University Press, 1998. [Google Scholar]
- Craddock, T. J.; Hameroff, S.; Ayoub, A. T.; Klobukowski, M.; Tuszynski, J. A. Anesthetics act in quantum channels in brain microtubules to prevent consciousness. Current Topics in Medicinal Chemistry 2015, 15(6), 523–533. [Google Scholar] [CrossRef]
- Davies, P. C. Does quantum mechanics play a non-trivial role in life? Biosystems 2004, 78(1-3), 69–79. [Google Scholar] [CrossRef]
- Dehaene, S. Consciousness and the Brain: Deciphering How the Brain Codes Our Thoughts; Viking, 2014. [Google Scholar]
- Del Giudice, E.; Preparata, G.; Vitiello, G. Water as a free electric dipole laser. Physical Review Letters 2005, 61(9), 1085–1088. [Google Scholar] [CrossRef]
- Dennett, D. C. Consciousness Explained; Little, Brown and Company, 1991. [Google Scholar]
- Dill, K. A.; MacCallum, J. L. The protein-folding problem, 50 years on. Science 2012, 338(6110), 1042–1046. [Google Scholar] [CrossRef]
- Edelman, G. M. Neural Darwinism: The Theory of Neuronal Group Selection; Basic Books, 1987. [Google Scholar]
- Edelman, G. M.; Tononi, G. A Universe of Consciousness: How Matter Becomes Imagination; Basic Books, 2000. [Google Scholar]
- Einevoll, G. T.; Kayser, C.; Logothetis, N. K.; Panzeri, S. Modelling and analysis of local field potentials for studying the function of cortical circuits. Nature Reviews Neuroscience 2013, 14(11), 770–785. [Google Scholar] [CrossRef]
- Engel, A. K.; Singer, W. Temporal binding and the neural correlates of sensory awareness. Trends in Cognitive Sciences 2001, 5(1), 16–25. [Google Scholar] [CrossRef]
- Engel, G. S.; Calhoun, T. R.; Read, E. L.; Ahn, T. K.; Mančal, T.; Cheng, Y. C.; Fleming, G. R. Evidence for wavelike energy transfer through quantum coherence in photosynthetic systems. Nature 2007, 446(7137), 782–786. [Google Scholar] [CrossRef]
- Erwin, D. H.; Valentine, J. W. The Cambrian Explosion: The Construction of Animal Biodiversity; Roberts and Company Publishers, 2013. [Google Scholar]
- Farris, S. M. Insect societies and the social brain. Current Opinion in Insect Science 2016, 15, 1–8. [Google Scholar] [CrossRef]
- Feinberg, T. E.; Mallatt, J. M. The Ancient Origins of Consciousness: How the Brain Created Experience; MIT Press, 2016. [Google Scholar]
- Fernald, R. D. Casting a genetic light on the evolution of eyes. Science 2006, 313(5795), 1914–1918. [Google Scholar] [CrossRef]
- Fortey, R.; Owens, R. M. Feeding habits in trilobites. Palaeontology 1999, 42(3), 429–465. [Google Scholar] [CrossRef]
- Franklin, R. E.; Gosling, R. G. Molecular configuration in sodium thymonucleate. Nature 1953, 171(4356), 740–741. [Google Scholar] [CrossRef]
- Freeman, W. J. Origin, structure, and role of background EEG activity. Part 1. Analytic amplitude. Clinical Neurophysiology 2004, 115(9), 2077–2088. [Google Scholar] [CrossRef] [PubMed]
- Freeland, S. J.; Hurst, L. D. The genetic code is one in a million. Journal of Molecular Evolution 1998, 47(3), 238–248. [Google Scholar] [CrossRef] [PubMed]
- Friedberg, E. C.; Walker, G. C.; Siede, W.; Wood, R. D.; Schultz, R. A.; Ellenberger, T. DNA Repair and Mutagenesis; ASM Press, 2006. [Google Scholar]
- Fröhlich, H. Theoretical physics and biology. In Biological Coherence and Response to External Stimuli; Springer, 1988; pp. 1–24. [Google Scholar]
- Gal, J.; Horvath, G.; Clarkson, E. N.; Haiman, O. Image formation by bifocal lenses in a trilobite eye? Vision Research 2000, 40(7), 843–853. [Google Scholar] [CrossRef] [PubMed]
- George, M. S.; Lisanby, S. H.; Avery, D.; McDonald, W. M.; Durkalski, V.; Pavlicova, M.; Sackeim, H. A. Daily left prefrontal transcranial magnetic stimulation therapy for major depressive disorder: a sham-controlled randomized trial. Archives of General Psychiatry 2010, 67(5), 507–516. [Google Scholar] [CrossRef]
- Hagan, S.; Hameroff, S. R.; Tuszynski, J. A. Quantum computation in brain microtubules: decoherence and biological feasibility. Physical Review E 2002, 65(6), 061901. [Google Scholar] [CrossRef]
- Hallett, M. Transcranial magnetic stimulation: a primer. Neuron 2007, 55(2), 187–199. [Google Scholar] [CrossRef]
- Hameroff, S. Consciousness, neurobiology and quantum mechanics: the case for a connection. In The Emerging Physics of Consciousness; Springer, 2006; pp. 193–253. [Google Scholar]
- Hameroff, S.; Penrose, R. Consciousness in the universe: a review of the 'Orch OR' theory. Physics of Life Reviews 2014, 11(1), 39–78. [Google Scholar] [CrossRef]
- Hille, B. Ion Channels of Excitable Membranes; Sinauer Associates, 2001. [Google Scholar]
- Ho, M. W. The Rainbow and the Worm: The Physics of Organisms; World Scientific, 1993. [Google Scholar]
- Ho, M. W. Bioenergetics and the coherence of organisms. NeuroNetwork World 1995, 5, 733–750. [Google Scholar]
- Ho, M. W. The Rainbow and the Worm: The Physics of Organisms, 2nd ed.; World Scientific Publishing, 1998. [Google Scholar]
- Ho, M. W. The Rainbow and the Worm: The Physics of Organisms, 3rd ed.; World Scientific Publishing, 2008. [Google Scholar]
- Ho, M. W.; Knight, D. P. The actin cytoskeleton in living cells is a dynamic coherent liquid crystal. BioSystems 1998, 46(1-2), 41–55. [Google Scholar]
- Hodgkin, A. L.; Huxley, A. F. A quantitative description of membrane current and its application to conduction and excitation in nerve. Journal of Physiology 1952, 117(4), 500–544. [Google Scholar] [CrossRef] [PubMed]
- Holland, N. D. Early central nervous system evolution: an era of skin brains? Nature Reviews Neuroscience 2003, 4(8), 617–627. [Google Scholar] [CrossRef] [PubMed]
- Hore, P. J.; Mouritsen, H. The radical-pair mechanism of magnetoreception. Annual Review of Biophysics 2016, 45, 299–344. [Google Scholar] [CrossRef]
- Hubel, D. H.; Wiesel, T. N. Receptive fields, binocular interaction and functional architecture in the cat's visual cortex. Journal of Physiology 1962, 160(1), 106–154. [Google Scholar] [CrossRef]
- Hunt, T. Neurotherapy applications of electromagnetic field theories of consciousness. Frontiers in Human Neuroscience 2024, 18, 1248424. [Google Scholar]
- Hunt, T.; Jones, M. Fields or firings? Comparing the spike code and the electromagnetic field hypothesis. Frontiers in Psychology 2023, 14, 1029715. [Google Scholar] [CrossRef]
- John, E. R. The neurophysics of consciousness. Brain Research Reviews 2002, 39(1), 1–28. [Google Scholar] [CrossRef]
- Joyce, G. F. The antiquity of RNA-based evolution. Nature 2002, 418(6894), 214–221. [Google Scholar] [CrossRef]
- Kamerlin, S. C.; Warshel, A. At the dawn of the 21st century: is dynamics the missing link for understanding enzyme catalysis? Proteins 2010, 78(6), 1339–1375. [Google Scholar] [CrossRef] [PubMed]
- Kandel, E. R. The molecular biology of memory storage: a dialogue between genes and synapses. Science 2001, 294(5544), 1030–1038. [Google Scholar] [CrossRef] [PubMed]
- Kandel, E. R.; Schwartz, J. H.; Jessell, T. M.; Siegelbaum, S. A.; Hudspeth, A. J. Principles of Neural Science, 5th ed.; McGraw-Hill, 2012. [Google Scholar]
- Katz, B.; Miledi, R. Further study of the role of calcium in synaptic transmission. Journal of Physiology 1970, 207(3), 789–801. [Google Scholar] [CrossRef] [PubMed]
- Katz, B.; Schmitt, O. H. Electric interaction between two adjacent nerve fibres. Journal of Physiology 1940, 97(4), 471–488. [Google Scholar] [CrossRef]
- Koonin, E. V.; Martin, W. On the origin of genomes and cells within the RNA world. Genome Biology 2005, 6(4), 214. [Google Scholar]
- Lambert, N.; Chen, Y. N.; Cheng, Y. C.; Li, C. M.; Chen, G. Y.; Nori, F. Quantum biology. Nature Physics 2013, 9(1), 10–18. [Google Scholar] [CrossRef]
- Land, M. F.; Nilsson, D. E. Animal Eyes; Oxford University Press, 2012. [Google Scholar]
- Lane, N. The Vital Question: Energy, Evolution, and the Origins of Complex Life; W. W. Norton & Company, 2015. [Google Scholar]
- Lane, N.; Martin, W. The energetics of genome complexity. Nature 2010, 467(7318), 929–934. [Google Scholar] [CrossRef]
- Laughlin, S. B.; de Ruyter van Steveninck, R. R.; Anderson, J. C. The metabolic cost of neural information. Nature Neuroscience 1998, 1(1), 36–41. [Google Scholar] [CrossRef]
- Lazcano, A.; Guerrero, R.; Margulis, L.; Oró, J. The evolutionary transition from RNA to DNA in early cells. Journal of Molecular Evolution 1988, 27(4), 283–290. [Google Scholar] [CrossRef]
- Lee, M. S.; Budney, L. A.; Clarkson, E. N. Compound eye evolution: highly conserved retinula and cone cell patterns indicate common origin of lithofacies eyes in trilobites. Evolution & Development 2007, 9(2), 155–164. [Google Scholar]
- Lennie, P. The cost of cortical computation. Current Biology 2003, 13(6), 493–497. [Google Scholar] [CrossRef]
- Levin, M.; Pezzulo, G.; Finkelstein, J. M. Endogenous bioelectric signaling networks: exploiting voltage gradients for control of growth and form. Annual Review of Biomedical Engineering 2017, 19, 353–387. [Google Scholar] [CrossRef]
- Lindahl, T. Instability and decay of the primary structure of DNA. Nature 1993, 362(6422), 709–715. [Google Scholar] [CrossRef]
- Lisman, J. E.; Jensen, O. The theta-gamma neural code. Neuron 2013, 77(6), 1002–1016. [Google Scholar] [CrossRef]
- Livingstone, M.; Hubel, D. Segregation of form, color, movement, and depth: anatomy, physiology, and perception. Science 1988, 240(4853), 740–749. [Google Scholar] [CrossRef] [PubMed]
- Llinas, R.; Ribary, U.; Contreras, D.; Pedroarena, C. The neuronal basis for consciousness. Philosophical Transactions of the Royal Society B 1998, 353(1377), 1841–1849. [Google Scholar]
- Marais, A.; Adams, B.; Ringsmuth, A. K.; Ferretti, M.; Gruber, J. M.; Hendrikx, R.; van Grondelle, R. The future of quantum biology. Journal of the Royal Society Interface 2018, 15(148), 20180640. [Google Scholar] [CrossRef] [PubMed]
- Marshall, C. R. Explaining the Cambrian "explosion" of animals. Annual Review of Earth and Planetary Sciences 2006, 34, 355–384. [Google Scholar] [CrossRef]
- Marshall, C. R.; Ahlberg, P. E. Cambrian stem-group arthropods and the case for a solution to the 'small shelly fossils' problem. Proceedings of the Royal Society B 2014, 281(1779), 20132716. [Google Scholar]
- Martin, S. J.; Grimwood, P. D.; Morris, R. G. Synaptic plasticity and memory: an evaluation of the hypothesis. Annual Review of Neuroscience 2000, 23(1), 649–711. [Google Scholar] [CrossRef]
- Mathews, J.; Levin, M. The body electric 2.0: recent advances in developmental bioelectricity for regenerative and synthetic bioengineering. Current Opinion in Biotechnology 2018, 52, 134–144. [Google Scholar] [CrossRef]
- McCaig, C. D.; Rajnicek, A. M.; Song, B.; Zhao, M. Controlling cell behavior electrically: current views and future potential. Physiological Reviews 2005, 85(3), 943–978. [Google Scholar] [CrossRef] [PubMed]
- McFadden, J. Integrating information in the brain's EM field: the cemi field theory of consciousness. Neuroscience of Consciousness 2020, 2020(1), niaa016. [Google Scholar] [CrossRef] [PubMed]
- McFadden, J.; Al-Khalili, J. The origins of quantum biology. Proceedings of the Royal Society A 2018, 474(2220). [Google Scholar] [CrossRef]
- Merker, B. The liabilities of mobility: a selection pressure for the transition to consciousness in animal evolution. Consciousness and Cognition 2005, 14(1), 89–114. [Google Scholar] [CrossRef] [PubMed]
- Mitchell, P. Coupling of phosphorylation to electron and hydrogen transfer by a chemi-osmotic type of mechanism. Nature 1961, 191(4784), 144–148. [Google Scholar] [CrossRef]
- Moroz, L. L. On the independent origins of complex brains and neurons. Brain, Behavior and Evolution 2009, 74(3), 177–190. [Google Scholar] [CrossRef]
- Mouritsen, H.; Ritz, T. Magnetoreception and its use in bird navigation. Current Opinion in Neurobiology 2005, 15(4), 406–414. [Google Scholar] [CrossRef]
- Nicholls, D. G.; Ferguson, S. J. Bioenergetics; Academic Press, 2013. [Google Scholar]
- Nilsson, D. E. Eye evolution: a question of genetic promiscuity. Current Opinion in Neurobiology 2004, 14(4), 407–414. [Google Scholar] [CrossRef]
- Nilsson, D. E. Eye evolution and its functional basis. Visual Neuroscience 2013, 30(1-2), 5–20. [Google Scholar] [CrossRef] [PubMed]
- Nilsson, D. E.; Pelger, S. A pessimistic estimate of the time required for an eye to evolve. Proceedings of the Royal Society B 1994, 256(1345), 53–58. [Google Scholar] [PubMed]
- Nitsche, M. A.; Cohen, L. G.; Wassermann, E. M.; Priori, A.; Lang, N.; Antal, A.; Pascual-Leone, A. Transcranial direct current stimulation: state of the art 2008. Brain Stimulation 2008, 1(3), 206–223. [Google Scholar] [CrossRef]
- Northcutt, R. G. Evolution of centralized nervous systems: two schools of evolutionary thought. Proceedings of the National Academy of Sciences 2012, 109 (Supplement 1), 10626–10633. [Google Scholar] [CrossRef] [PubMed]
- Packard, A. Cephalopods and fish: the limits of convergence. Biological Reviews 1972, 47(2), 241–307. [Google Scholar] [CrossRef]
- Pai, V. P.; Aw, S.; Shomrat, T.; Lemire, J. M.; Levin, M. Transmembrane voltage potential controls embryonic eye patterning in Xenopus laevis. Development 2012, 139(2), 313–323. [Google Scholar] [CrossRef]
- Panitchayangkoon, G.; Hayes, D.; Fransted, K. A.; Caram, J. R.; Harel, E.; Wen, J.; Engel, G. S. Long-lived quantum coherence in photosynthetic complexes at physiological temperature. Proceedings of the National Academy of Sciences 2010, 107(29), 12766–12770. [Google Scholar] [CrossRef]
- Parker, A. R. the Blink of an Eye: How Vision Sparked the Big Bang of Evolution; Perseus Publishing, 2003. [Google Scholar]
- Parker, A. R. The Optics of Life: A Biologist's Guide to Light in Nature; Princeton University Press, 2011. [Google Scholar]
- Paterson, J. R.; García-Bellido, D. C.; Lee, M. S.; Brock, G. A.; Jago, J. B.; Edgecombe, G. D. Acute vision in the giant Cambrian predator Anomalocaris and the origin of compound eyes. Nature 2011, 480(7376), 237–240. [Google Scholar] [CrossRef]
- Pauling, L. The Nature of the Chemical Bond; Cornell University Press, 1960. [Google Scholar]
- Penrose, R. The Emperor's New Mind: Concerning Computers, Minds, and the Laws of Physics; Oxford University Press, 1989. [Google Scholar]
- Pinotsis, D. A.; Miller, E. K. In vivo ephaptic coupling allows memory network formation. Cerebral Cortex 2023, 33(10), 6000–6007. [Google Scholar] [CrossRef]
- Pockett, S. The Nature of Consciousness: A Hypothesis; Writers Club Press, 2000. [Google Scholar]
- Poole, A. M.; Jeffares, D. C.; Penny, D. The path from the RNA world. Journal of Molecular Evolution 1998, 46(1), 1–17. [Google Scholar] [CrossRef]
- Popp, F. A.; Li, K. H.; Gu, Q. (Eds.) Recent Advances in Biophoton Research and Its Applications; World Scientific, 1992. [Google Scholar]
- Radman, T.; Su, Y.; An, J. H.; Parra, L. C.; Bikson, M. Spike timing amplifies the effect of electric fields on neurons: implications for endogenous field effects. Journal of Neuroscience 2007, 27(11), 3030–3036. [Google Scholar] [CrossRef]
- Reddien, P. W.; Alvarado, A. S. Fundamentals of planarian regeneration. Annual Review of Cell and Developmental Biology 2004, 20, 725–757. [Google Scholar] [CrossRef] [PubMed]
- Revonsuo, A.; Newman, J. Binding and consciousness. Consciousness and Cognition 1999, 8(2), 123–127. [Google Scholar] [CrossRef] [PubMed]
- Ritz, T.; Adem, S.; Schulten, K. A model for photoreceptor-based magnetoreception in birds. Biophysical Journal 2000, 78(2), 707–718. [Google Scholar] [CrossRef]
- Rossi, S.; Hallett, M.; Rossini, P. M.; Pascual-Leone, A. Safety, ethical considerations, and application guidelines for the use of transcranial magnetic stimulation in clinical practice and research. Clinical Neurophysiology 2009, 120(12), 2008–2039. [Google Scholar] [CrossRef]
- Ruffini, G.; Wendling, F.; Sanchez-Todo, R.; Santarnecchi, E. Targeting brain networks with multichannel transcranial current stimulation (tCS). Current Opinion in Biomedical Engineering 2020, 8, 70–77. [Google Scholar] [CrossRef]
- Sagan, C. Ultraviolet selection pressure on the earliest organisms. Journal of Theoretical Biology 1973, 39(1), 195–200. [Google Scholar] [CrossRef]
- Saenger, W. Principles of Nucleic Acid Structure; Springer, 1984. [Google Scholar]
- Sancar, A. DNA excision repair. Annual Review of Biochemistry 1996, 65(1), 43–81. [Google Scholar] [CrossRef]
- Schoenemann, B.; Clarkson, E. N. Discovery of some 400 million year-old sensory structures in the compound eyes of trilobites. Scientific Reports 2013, 3, 1429. [Google Scholar] [CrossRef]
- Shannon, C. E. A mathematical theory of communication. Bell System Technical Journal 1948, 27(3), 379–423. [Google Scholar] [CrossRef]
- Still, S.; Sivak, D. A.; Bell, A. J.; Crooks, G. E. Thermodynamics of prediction. Physical Review Letters 2012, 109(12), 120604. [Google Scholar] [CrossRef] [PubMed]
- Stock, D.; Leslie, A. G.; Walker, J. E. Molecular architecture of the rotary motor in ATP synthase. Science 1999, 286(5445), 1700–1705. [Google Scholar] [CrossRef] [PubMed]
- Strausfeld, N. J.; Hirth, F. Deep homology of arthropod central complex and vertebrate basal ganglia. Science 2013, 340(6129), 157–161. [Google Scholar] [CrossRef] [PubMed]
- Tegmark, M. Importance of quantum decoherence in brain processes. Physical Review E 2000, 61(4), 4194–4206. [Google Scholar] [CrossRef]
- Tomer, R.; Denes, A. S.; Tessmar-Raible, K.; Arendt, D. Profiling by image registration reveals common origin of annelid mushroom bodies and vertebrate pallium. Cell 2010, 142(5), 800–809. [Google Scholar] [CrossRef]
- Tononi, G. An information integration theory of consciousness. BMC Neuroscience 2004, 5(1), 42. [Google Scholar] [CrossRef]
- Treisman, A. The binding problem. Current Opinion in Neurobiology 1996, 6(2), 171–178. [Google Scholar] [CrossRef]
- Tsodyks, M. V. Associative memory in neural networks with the Hebbian learning rule. Modern Physics Letters B 1999, 3(8), 555–560. [Google Scholar] [CrossRef]
- VanRullen, R.; Koch, C. Is perception discrete or continuous? Trends in Cognitive Sciences 2003, 7(5), 207–213. [Google Scholar] [CrossRef]
- Walker, J. E. The ATP synthase: the understood, the uncertain and the unknown. Biochemical Society Transactions 2013, 41(1), 1–16. [Google Scholar] [CrossRef]
- Walker, S. I.; Davies, P. C.; Ellis, G. F. From Matter to Life: Information and Causality; Cambridge University Press, 2017. [Google Scholar]
- Wang, X. J. Neurophysiological and computational principles of cortical rhythms in cognition. Physiological Reviews 2010, 90(3), 1195–1268. [Google Scholar] [CrossRef]
- Ward, L. M. Synchronous neural oscillations and cognitive processes. Trends in Cognitive Sciences 2003, 7(12), 553–559. [Google Scholar] [CrossRef]
- Warshel, A.; Sharma, P. K.; Kato, M.; Xiang, Y.; Liu, H.; Olsson, M. H. Electrostatic basis for enzyme catalysis. Chemical Reviews 2006, 106(8), 3210–3235. [Google Scholar] [CrossRef] [PubMed]
- Watson, J. D.; Crick, F. H. Molecular structure of nucleic acids. Nature 1953, 171(4356), 737–738. [Google Scholar] [CrossRef]
- Weinberg, S. The Quantum Theory of Fields; Cambridge University Press, 1995. [Google Scholar]
- Yoshida, M.; Muneyuki, E.; Hisabori, T. ATP synthase—a marvellous rotary engine of the cell. Nature Reviews Molecular Cell Biology 2001, 2(9), 669–677. [Google Scholar] [CrossRef] [PubMed]
- Young, J. A.; Lee, S.; Kumar, R. Computational power and cephalization in early bilaterian evolution. Journal of Evolutionary Biology 2024, 37(3), 245–258. [Google Scholar]
- Zhao, M. Electrical fields in wound healing—an overriding signal that directs cell migration. Seminars in Cell & Developmental Biology 2009, 20(6), 674–682. [Google Scholar] [CrossRef]
- Brake, N.; Duc, F.; Rokos, A.; Arseneau, F.; Shahiri, S.; Khadra, A.; Plourde, G. A neurophysiological basis for aperiodic EEG and the background spectral trend. Nature Communications 2024, 15(1), 1514. [Google Scholar] [CrossRef]
- Buzsaki, G. Rhythms of the Brain; Oxford University Press, 2006. [Google Scholar]
- Canolty, R. T.; Knight, R. T. The functional role of cross-frequency coupling. Trends in Cognitive Sciences 2010, 14(11), 506–515. [Google Scholar] [CrossRef]
- Dennett, D. C. Consciousness Explained; Little, Brown and Company, 1991. [Google Scholar]
- Edelman, G. M. Neural Darwinism: The Theory of Neuronal Group Selection; Basic Books, 1987. [Google Scholar]
- Feinberg, T. E.; Mallatt, J. M. The Ancient Origins of Consciousness: How the Brain Created Experience; MIT Press, 2016. [Google Scholar]
- Finley, A. J.; Angus, D. J.; Knight, E. L.; van Reekum, C. M.; Lachman, M. E.; Davidson, R. J.; Schaefer, S. M. Resting EEG periodic and aperiodic components predict cognitive decline over 10 years. Journal of Neuroscience 2024, 44(13), e1332232024. [Google Scholar] [CrossRef]
- Francis, J. T.; Gluckman, B. J.; Schiff, S. J. Sensitivity of neurons to weak electric fields. Journal of Neuroscience 2003, 23(19), 7255–7261. [Google Scholar] [CrossRef]
- Freeman, W. J. The physiology of perception. Scientific American 1991, 264(2), 78–85. [Google Scholar] [CrossRef]
- Freeman, W. J. Origin, structure, and role of background EEG activity. Part 1. Analytic amplitude. Clinical Neurophysiology 2004, 115(9), 2077–2088. [Google Scholar] [CrossRef]
- Hales, C. G.; Ericson, M. Consciousness: The electromagnetic turn. Frontiers in Human Neuroscience 2022, 16, 836046. [Google Scholar]
- John, E. R. The neurophysics of consciousness. Brain Research Reviews 2002, 39(1), 1–28. [Google Scholar] [CrossRef]
- Merker, B. The liabilities of mobility: a selection pressure for the transition to consciousness in animal evolution. Consciousness and Cognition 2005, 14(1), 89–114. [Google Scholar] [CrossRef]
- Revonsuo, A.; Newman, J. Binding and consciousness. Consciousness and Cognition 1999, 8(2), 123–127. [Google Scholar] [CrossRef]
- Treisman, A. The binding problem. Current Opinion in Neurobiology 1996, 6(2), 171–178. [Google Scholar] [CrossRef] [PubMed]
- VanRullen, R.; Koch, C. Is perception discrete or continuous? Trends in Cognitive Sciences 2003, 7(5), 207–213. [Google Scholar] [CrossRef] [PubMed]
- Ward, L. M. Synchronous neural oscillations and cognitive processes. Trends in Cognitive Sciences 2003, 7(12), 553–559. [Google Scholar] [CrossRef] [PubMed]
|
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. |
© 2025 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access article distributed under the terms and conditions of the Creative Commons Attribution (CC BY) license (http://creativecommons.org/licenses/by/4.0/).