2.1. Resuming Some of the Problems within the Corpuscular QM View (1st Quantization)
As a matter of fact, the whole picture on which Pauling, the IUPAC, and the most of chemists, based their stances, counts just for describing matter through a corpuscular approach (both classical and Quantum Mechanical, QM) which is not able to explain the tricky features of condensed systems like liquid water. Indeed, the problem is that, according to Maxwell’s equations (Langevin’s theory), liquid water should have a relative dielectric constant given by:
A first problem is that, with NA the Avogadro number, p0 = 1.85498 D (D: Debye, such as 1 D = 3.336·10−30 C·m), ρ = 1 g·cm-3 and M = 18 g·mol-1, we predict εr (T = 300 K) ≈ 13 instead of the experimental value of εr = 80 for bulk water. Moreover, this approach is not able to explain why the solvent power of water for ionic salt increases with temperature while at the same time the dielectric permittivity diminishes (Del Giudice & Preparata, 2000). In order to yield the electrical rigidity of water molecules (with polarizability volume α’ = 1.47 Å3, ionization potential, or ionization threshold, Ith = 12.6 eV, molecular radius R =v2.4 Å), Maxwell equations suggest dipolar interaction energies (Matcha & King, 1976) equal to:
These calculations assume an oxygen-oxygen distance of 3.65 Å, corresponding to the O-H covalent bond length (0.95 Å), augmented by the sum of van der Waals radii of hydrogen (1.2 Å) and oxygen (1.5 Å). Such values cannot explain the abnormally high boiling point of liquid water, or the HB energy of about 0.2 eV. Moreover, not even summing these classical-theory values does help, since the HBs’ energy depends on the molecular environment of the water molecules. Indeed HB energy is said, for instance, to be 0.15 eV in the “water dimer”, 0.24 eV in liquid water and 0.29 eV in hexagonal ice (Ghanty, Staroverov, Koren, & Davidson, 2000) and its apparent covalency puts in doubt its electrostatic nature (Isaacs, et al., 1999). The problem is that it’s impossible to measure each classical/QM interaction (Keesom, Debye and London) separately, they can only be evaluated by using a theoretical model and these interactions would remain exactly the same whatever the aggregation state of the water molecules is. So, from an energetic viewpoint, HBs behave quite differently from van der Waals forces (Keesom, Debye, London) (Henry M. , 2014).
Especially for liquid water, the agreement of theoretical models – based on the QM picture of matter – to the experimental results and the well-known water anomalies is not satisfying at all. In this regard we quote Marc Henry (Henry M. , 2015):
«The claim that chemistry has been completely explained in terms of quantum theory is now received wisdom among physicists and chemists. Yet quantum physics is able neither to predict nor explain the strong association of water molecules in liquid or ice. Quantum chemistry algorithms either exclude hydrogen bonded (H-bonded) systems, or treat them by modelling a water molecule as an asymmetric tetrahedron having two positive and two negative electrical charges at its vertices. Recent calculations of the potential energy surface of the simple water dimer {H2O}2 yield 30,000 ab initio energies at the coupled clustering techniques (CCT) level (Shank, Wang, Kaledin, Braams, & Bowman, 2009). But free OH-stretches [deviate from] experimental values by 30-40 cm-1 and their dissociation energy 1.1 kJ·mol-1 [are likewise] below benchmark experimental values. To obtain satisfactory agreement with experiment, it is necessary to replace ab initio potentials with spectroscopically accurate measurements. This is hardly a ringing endorsement of the underlying theory» (despite Dirac’s 1929 claims (Dirac, 1929)).
Indeed, Molecular Orbital (MO) Theory (basically an application of a QM-first quantisation theory to the molecular orbital approximations) studies molecular bonding by approximating the positions of bonded electrons through a Linear Combination of their Atomic Orbitals (LCAO). This is achieved, for example, by applying the Hartree-Fock model to Schrödinger’s equation (Matcha & King, 1976). But in the LCAO picture there are big problems about the topology and overlap of orbitals of the water molecule. If we consider the basic C2v-symmetry of H2O (according to the Schönflies classification (Flurry, 1980)), water has four irreducible representations named a1, a2, b1 and b2, where “a” (“b”) indicates symmetric (anti-symmetric) representation with respect to a rotation around the main symmetry axis, in this case the z-axis, the same one along which the pz oxygen orbital is oriented. The subscripts “1” and “2”, respectively, indicate symmetric and antisymmetric representations with respect to the rotation around a C2 axis, perpendicular to the main symmetry axis, or with respect to a plane σv, if C2 is missing.
In a single water molecule, we have ten electrons (their occupation number for each orbital is given by the superscript out of the brackets in the following expression) that must be distributed among five energy levels according to the electronic configuration: (1a
1)
2(2a
1)
2(1b
2)
2(3a
1)
2(1b
1)
2(4a
1)
0(2b
2)
0 (see
Figure 1).
Accordingly, this does not allow the establishment of the partial covalence involving the Highest Occupied Molecular Orbital (HOMO), displaying b1-symmetry (non-symmetric with respect to the z axis, the same along which is oriented the pz oxygen orbital) and the Lowest Unoccupied Molecular Orbital (LUMO), displaying a1-symmetry (symmetric with respect to the z axis). Thus, no HOMO-LUMO interaction overlap can possibly occur.
Even by arguing that, during hydrogen bonding, the symmetry is lowered, thus leading to a possible non-zero overlap, is still unsatisfying because prior to the HB, both partners display their full C
2v-symmetry with zero overlap, while, from the experiments, we know that the final symmetry of water dimers, or more numerous aggregates, is C
s (reflection with respect to a
σ-plane). So, at what distance would the symmetry change from C
2v to C
s? The assumption that C
s-symmetry would be held at every distance is useless because the HOMO level would still represent one symmetry, and the LUMO another: the overlapping integral would be again zero. Someone could think that overlaps may occur through other molecular orbitals, describing the covalent O-H as
σ-bonds, leaving two outer non-equivalent “lone-pairs” (3a
1, 1b
1) available for making HB with other water molecules, but at both 2.75 Å (the distance reached by 3a
1) and 2.98 Å (the distance reached by 1b
1), the overlap between the acceptor oxygen and the hydrogen-bonding proton is negative, because the 3a
1 (HOMO-1)
1 and the 1b
1 (HOMO) have two very different topologies and energies, pointing to a net anti-bonding covalent interaction in the quantum sense (Ghanty, Staroverov, Koren, & Davidson, 2000). Furthermore, X-ray emission spectroscopy (XES) reveals that in a water molecule the 1b
1 HOMO-level is not affected by the HB [23], while, on the contrary, a strong perturbation of the 3a
1 (HOMO-1) level is observed. This is evidence for a rather unconventional (within the LCAO QM picture) HOMO-1/LUMO interaction (see
Figure 2). In addition, Compton scattering experiments which revealed a strong anisotropy of the momentum density of valence electrons) in hexagonal ice (I
h-type), is evidence of a neat anti-bonding, repulsive, interaction between neighbouring water molecules despite the multicentred character of the QM wave functions (Romero, Silvestrelli, & Parrinello, 2001). Eventually, topological analysis of electronic density revealed that it’s not possible to distinguish between HBs and mere van der Waals interactions (Bader, 1990).
Given the experimental photoelectron spectrum of the water molecule (
Figure 3), its most faithful representation should display three kinds of orbitals (two σ-bonds, one 2s-type lone pair and one 2p-type lone pair), and not two types (two σ-bonds and two equivalent lone pairs), as suggested by MO theory (Becke & Edgecombe, 1990). The only way to retrieve a physical picture involving two lone pairs and two σ-bonds approximately oriented towards the vertices of a tetrahedron, is to look at the positions of the largest eigenvalues and corresponding eigenvectors of the Hessian minima in their molecular electrostatic potential (Kumar, Gadre, Mohan, & Suresh, 2014). But, again, this means reverting to a purely electrostatic view of HBs with all the annexed issues we’re listing out.
The situation is so confusing that the scientific community today is divided into two opposing camps, one promoting water as a random tetrahedral network with flickering HBs (Bukowski, Szalewicz, Groenenboom, & van der Avoird, 2007), and the other promoting water in terms of a two-state model, one being tetrahedral, the other not (Wernet, et al., 2004) [23], despite a very different picture is obtained if X-rays are absorbed by water, as the water appears not arranged in a local tetrahedral geometry but rather formed of entangled chains (Wernet P. e., 2004). However, this two-state model is closer to the truth, but the underlying idea on which it is based – still corpuscular QM – implies a physically unmotivated “cut-off energy” at which the two populations of molecules should be separated (this point will be elaborated further below). By picturing liquid water as a flickering network of HBs, the problem lays in how this interaction is commonly conceived and treated in the theoretical models: electrostatically or electrodynamically but still in a perturbative way and in a first quantisation only. This is not satisfying, indeed, neutron scattering experiments
2, as well as molecular dynamic simulations, have shown that the average residence-time of a hydrogen atom around a water molecule is, in average, about 2 ps at T = 300 K, and increases to 20 ps at T = 250 K (Teixeira, Bellissent-Funel, Chen, & Dianoux, 1985). Electrical charges moving on the picosecond timescale, may be expected to generate an electromagnetic field with at least a frequency of the order of 10
12 Hz. However, the electromagnetic fields should not be treated classically, instead the interaction of molecules with the ubiquitous vacuum electromagnetic fluctuations must be considered to properly understand the condensation (Preparata G. , 1995).
-
1HOMO-1, HOMO-2, HOMO-3, …, HOMO-N or LUMO+1, LUMO+2, LUMO+3, …, LUMO+N denote electronic levels, among the several molecular orbitals, placed at the Nth level below (-) the HOMO or at the Nth level above (+) the LUMO.
-
2The problem with X-ray scattering, indeed, is that it tends to give a static image of water, whereas it is a dynamic medium. Neutron scattering, on the contrary, has revealed the existence of two relaxation times in liquid water (Teixeira, Bellissent-Funel, Chen, & Dianoux, 1985). Thus, the first time close to 1-2 ps at room temperature corresponds to the fluctuation of the network of hydrogen bonds following the rotations of the water molecules. This relaxation time follows an Arrhenius law τLH= τ0 ∙ exp(U#/kBT) with τ0=0.0485 ps and an activation energy U# = 7.7 kJ/mol. As for the second relaxation time, it varies very strongly with temperature, from 1.25 ps at 20 °C to 22.7 ps at −20 °C. This indicates that two fraction exists, and one of the two has intrinsic dynamics independent on temperature (the one which is coherent, as it will be discussed deeply in the following).
2.1. Synthesis of the Theoretical Background in QFT-QED for Liquid Water
By adopting a QFT-QED description of liquid water (and condensed matter in general) where a complementarity relationship between phase of oscillation and number of oscillating quanta emerges, the HB is consistently countable for as an emergent property of boson condensation at a new (lower) ground level (vacuum) of water molecules. Under these circumstances their dipolar oscillations are kept in phase by a coupled and self-trapped electromagnetic (em) vector potential (A), as it has been shown by relevant literature (Arani, Bono, Del Giudice, & Preparata, 1995), (Bono, Del Giudice, Gamberale, & Henry, 2012), (Del Giudice, Galimberti, Gamberale, & Preparata, 1995), (Preparata G. , 1995).
By letting decay some approximations in the matter-em-field interaction, (like the SVE,
3 valid for a system of isolated particles and displaying a finite number of degrees of freedom), and by moving to the “2nd quantisation” of QFT (having an infinite number of degrees of freedom), new profound insights about liquid water and the nature of HB emerge. In the QFT-QED picture the cohesive energy emerges to come from the coherence energy gap (denoted as Δ
g or E
coh) associated to millions of water molecules being phase-locked and packed together by a self-trapped em-field and not from the sum of individual incoherent, directional, interactions as it’s usually deemed within the framework of the 1st quantization.
In doing so we briefly summarize the two-fluid picture for water as it emerges from the QED theory, firstly developed in the ‘80s, (Preparata, Del Giudice, & Vitiello, 1988) (Arani, Bono, Del Giudice, & Preparata, 1995) (Preparata G. , 1995) (Bono, Del Giudice, Gamberale, & Henry, 2012) and show some crucial experimental data which endorse this theoretical approach.
Despite the thermodynamics of water liquefaction from vapour has been precisely tackled in classical physical-chemistry (enabling quantifying the amounts of entropy variation and latent heat) (Franks, 1972-1982), within a purely QM corpuscular picture (which relies, for the condensation to a liquid state, on the establishment of a flickering network of local, directional, intermolecular forces, such as the deemed “HBs”), the real physical origin of such high values of entropy variation, boiling temperature and latent heat of vaporisation are not fully derivable.
The main limits in the framework of corpuscular QM consist mainly in two points interwoven with each other: (i) the inability to consider systems with large amounts of, or infinite, degrees of freedom (where the number operator , is left undefined) (Preparata G. , 2002), and (ii) the inability to describe symmetry breakings (i.e.,: phase transitions), being a theory which obeys the von Neumann’s theorem (von Neumann, 1955). According to this theorem, only one ground state, vacuum level, is considered, thus rendering unfeasible the description of symmetry breakings and phase transitions (Blasone, Jizba, & Vitiello, 2011). The main problem associated to such limitations is the impossibility to predict non-trivial solutions for the equations of motion starting from the perturbative ground state of the system (like vapour being cooled down). Therein, the vacuum fluctuations (able to excite the electron of water molecule) are counted for, but – being of the other of δ ≈ 1 ppm (i.e.,: the Lamb shift (Lamb & Retherford, 1947)) – are deemed as negligible. Actually, when the number of matter quanta (molecules) overcomes a critical threshold, this coupling (between vacuum virtual excitations and matter), becomes so meaningful as to change dramatically the system layout because it is not proportional just to N but to N√N (Arani, Bono, Del Giudice, & Preparata, 1995) (Bono, Del Giudice, Gamberale, & Henry, 2012). Moreover, in a QM picture, beyond the investigated transition, excitations over other levels in the spectrum of the molecule are not taken into account for the evolution of the system (Teixeira & Luzar, Physics of liquid water: Structure and dynamics, 1999). Thus, when several billions of molecules are coupled with em-quanta of the vacuum, they are back-reformed by this new emergent condition and a purely bottom-up description (based on the mere summation of the interaction calculated over few quanta) does not deliver a truthful picture (Henry M. , 2014).
In a QFT-QED perspective, when water vapour is at the liquefaction threshold (for instance, pressure P = 1 atm and temperature T = 373.15 K), water molecules are in constant dialogue with virtual quanta popping out of vacuum. Based on their energetic content (ΔE), these quanta can excite the electrons of water molecule at several levels. Of course, this process isolated molecules do not produce any permanent energetic gain and the excitation lasts – in accordance with the Heisenberg relationship – just a short time
. The interesting aspect, looking at water photoemission spectrum (see
Figure 3), is that the first possible transition has an energy of about 7.5 eV, and the other energetic levels are placed at >10 eV. This means that the spatial range of such excitations (the wavelength, λ, of virtual photons) is at least in the order of 100 nm (i.e.,: about one thousand times bigger than the water molecule itself!).
The numerical density of water vapour, at its boiling point (T = 373.15 K at P=1 atm), is around 2∙10
19 molecules⋅cm
-3, it means that an em-excitation, able to set a water molecule on a different electron configuration, includes within its own volume (V~λ
3) about 20’000 molecules. The more the density increases (by lowering T, for instance) and the more the probability grows that the photon released by a previously excited molecule – originally adsorbed from the vacuum – could then be re-absorbed by another one. At a critical density the photons of other molecules in the volume get involved in the same dynamics, until a sizeable em-field is established and self-trapped in an ensemble of water molecules which steadily grows. In doing so it keeps sucking in millions of molecules until the volume is filled. This saturation level is determined by short-range forces at which molecules are closely packed (intermolecular distance >3.1 Å, which is larger than the molecular radius >1 Å) resulting in an increase of the molecular size (Arani, Bono, Del Giudice, & Preparata, 1995). All photons set in phase among each other and with the molecules oscillating between the two electron levels. This new ordered state occurs because it is thermodynamically more favourable, provided that the system is open and could dissipate excess energy as heat (entropy). For liquid water, it has been computed that the energy difference, the energy gap,
Δg, is 0.16 ±0.05 eV (Bono, Del Giudice, Gamberale, & Henry, 2012). This is the origin of the high latent heat of liquefaction, where an excess energy with respect to the one received by the original vacuum, is given back to the environment. In water, QED calculations, using several possible candidate levels of its spectrum (as depicted in
Figure 3), showed that this probability becomes 100 % for the 5d level (at 12.07 eV above the ground state), above the density threshold ρ
c ≥ 0.32 g/cm
3. In the election of this level as the favourite one to settle a coherent excitation supporting the formation of the liquid phase there are also other parameters like: excitation energy or frequency (
ωq), coupling constant between em-field and oscillating charges (
g), photon mass renormalization term (
µr), oscillator strength (
fq), renormalized frequency (
ωr), energy gap (
Δg), mixing angle (α) (see Tables 1 and 2 in ref. (Bono, Del Giudice, Gamberale, & Henry, 2012) for details).
The original frequency, ωq, of the exciting em-field, which now spends part of its lifetime as excited molecules (and no more as a free field) become renormalized to a lower value, ωr, which is associated to the phase variation and locking between em- and matter-field. This renormalization implies that the field is made of quasi-particles – according of the Anderson-Higgs-Kibble mechanism have imaginary mass (negative squared mass) (Anderson, 1958) – and thus unable to propagate outside the region where the coupling with the matter-field is in force (Anderson, 1984). Such a region is named coherence domain (CD) and constitutes a self-generated sub-radiant cavity of the trapped fields (Del Giudice & Vitiello, 2006). It’s appropriate to speak of “matter-field” since the CD is an open system where a permanent crossover of molecules in and out of the CD occurs and the number (N) of matter quanta is undefined, allowing the phase () to be well-defined. This complies with the QFT complementarity relationship holding for phase and number operators, such that the higher the uncertainty () in terms of number, the better defined is the phase. Thus the “fundamental uncertainty relationship” is expressed as: (in natural units, where ℏ = c = kB = 1) (Preparata G. , 1995).
The expulsion of a large amount of entropic energy per molecule (which tells us a lot about the physical origin of latent heat of liquid water condensation) sets the coherent ensemble of molecules (namely, about 6 millions per CD) on a lower, desirable, ground state (vacuum level) whose energy difference with respect of the isolated molecules’ is denoted as the
energy gap,
Δg. This energy difference represents a crucial quantity of the system and expresses its thermodynamical stability, defining how much energy must be spent to liberate a molecule from a CD:
where
g is the coupling constant between em-field and matter,
µr the photon mass renormalization term, and
A0 the maximum amplitude of the em-field (i.e., the em-vector potential). The renormalized frequency,
ωr is defined as a function of the energy (i.e., frequency) of the exciting virtual photon (equal of the energy difference existing between sp
3 and 5d orbitals):
whereby the phase variation of the em-field,
is defined by the phase-factors time derivatives of the ground (
) and excited (
) states of the matter-field. The em-field is characterized by:
where α is the mixing angle between the two (fundamental, “0”, and excited, “q”) levels, so that 0<α<π/2. In these calculations the Natural Units system is used (where ℏ = c = k
B = 1, and the elementary electrical charge e = 0.302814). As shown in
Figure 4, the coherent state results to be a time-weighted average of the new ground state (at 90 %) and the excited one (10 %) where molecules assume an expanded shape (due to the larger 5d volume). This comes with two major implications: (i) the coherent fraction is less denser than the interstitial vapour-like incoherent one, and (ii) the re-arranged shape of the water molecules produces a physical explanation of the electron-cloud protrusions necessary for some tetrahedral arrangements observed in some water systems through suitable techniques (Tokushima, et al., 2008) (Huang, et al., 2009) (Taschin, Bartolini, Eramo, Righini, & Torre, 2013).
Of course, if the temperature were to approach 0 K, the whole system would become fully coherent: this occurs only at T <220 K (Garbelli, 2000). At higher temperatures the system is made up of two populations of molecules (those gathered in CDs,
Fcoh(T) and those incoherent, constituting the vapour-like phase, placed interstitially among the CDs,
Finc(T)) whose relative quantities obey a sum-rule:
Fcoh(T) + Finc(T) = 1. The higher the temperature is, the lower the coherent fraction becomes, enabling thermal agitation being able to gradually erode more molecules from the CDs’ periphery, reducing their diameter. At room conditions (T = 300 K),
Fcoh ~40 % the effective size of the CDs is about 60 nm (
Figure 5).
Furthermore, it’s noteworthy to mention that the coherent oscillations implies that 10 % of electrons in each CD stay very close to the ionisation threshold (placed at 12.62 eV, about 0.5 eV above the 5d orbital). Thus, with about six million molecules per CD, there are about 0.6∙106 quasi-free electrons per CD. These electrons circulate on the periphery of the CD, being subjected to a repulsive ponderomotive force (see Equation (10) in the following) and are coherent, therefore they cannot dissipate energy by thermal relaxation, nor friction (yielding closed supercurrents, said cold vortexes). The implications of these aspect are huge, but out of the scope of this work, for which we refer to (Del Giudice & Tedeschi, 2009) (Del Giudice, Spinetti, & Tedeschi, 2012) (Del Giudice, Voeikov, Tedeschi, & Vitiello, 2015) (Renati P. , 2020), (Madl & Renati, 2023).
The existence of incoherent fraction, above 273.15 K allows water to be a liquid, whereas below that temperature the mobility is too low, thus inducing crystallization. However, the global density,
ρ(T), is lower, right because
Fcoh(T) is increased with respect to 277.14 K (4 °C), at which water shows its maximum density (
Figure 5), more than how much the incoherent fraction increased its density by lowering T from 277.15 to 273.15 K. Only by applying such a two-fluid model is it possible to predict form
first principles the peculiar density trend of liquid water (Arani, Bono, Del Giudice, & Preparata, 1995) (Preparata G. , 1995) (Garbelli, 2000):
The density of coherent fraction does not depend on the temperature and has been calculated to be 0.92 g/cm3 on the basis of the wider shape given by the mixing angle weights over the coherent oscillation, ρn(T) is the density of the normal (incoherent) fraction (Finc(T) = 1 – Fcoh(T)). See (Arani, Bono, Del Giudice, & Preparata, 1995) and (Henry M. , 2014) for further details.
Recently some important experimental data showed both (i) the impossibility to explain some features within the corpuscular-QM picture and (ii) the necessity to contemplate, within the models, coherence as a key property to correctly fit the experimental data.
The first case refers to infrared (IR) and near IR (NIR) analysis of water or water solutions spectra (of O-H stretch mode range, IR, or of its first harmonic, NIR) taken at different temperatures (De Ninno, Del Giudice, Gamberale, & Castellano, 2014), (Renati, Kovacs, De Ninno, & Tsenkova, 2019) whose trends showed the clear existence of an isosbestic points that expresses the existence of two populations of molecules which depend reciprocally on T. This, of course, is not a novelty, but what it’s worth to look at are the resulting van’t Hoff plots (i.e.,: the Log (equilibrium constant of the passage from one population to the other) vs 1/T) is linear, revealing that (i) the energy difference between the two states does not depend on T and (ii) that its slope is in good agreement with the energy gap predicted by QED theory. Moreover, in (Renati, Kovacs, De Ninno, & Tsenkova, 2019) it has been shown how the plot of the logarithm of the ratios between the spectral intensity of one population (distinguished from the other one by the isosbestic point) with respect to the total, taken at each temperature, plotted as a function of log T yields a straight line. This accounts for a scale-free behaviour, revealing the underlying coherent dynamics for the demonstrated isomorphism existing between self-similar (fractal) topologies and squeezed quantum coherent states (Celeghini, De Martino, De Siena, Rasetti, & Vitiello, 1995) (Celeghini, Rasetti, & Vitiello, 1992) (Vitiello, 2009).
The second case deals with the fit of dielectric permittivity of pure water and electrolytes water solutions in the range 0.2-1.5 THz (De Ninno, Nikollari, Missori, & Frezza, 2020), (Nikollari, De Ninno, & Frezza, 2023). The fit to the experimental data requires a two-fluid Debye model that mimics the electrical permittivity (both for the real and the imaginary part). However, in order to be effective over the whole spectral range, it requires an additional linear term (ξω, where ξ ≈ 0.47 ps) to the imaginary part of the dielectric function. This fact has a profound physical meaning because implies the violation of Kramers-Kronig (KK) relations (Toll, 1956) within the time span ξ. The KK relations express the causal relation between the forcing field and the charge displacement. This tiny violation, within a time scale right of the order of magnitude of the renormalized oscillation period of the coherent field within the CDs (which excite and relax in a few hundreds of femtoseconds, τr ≈ 1/ωr ~ 300-500 fs) witnesses temporally non-local correlations in the medium (i.e., phase correlations), possible if the system is in an entangled coherent state (a phase eigenstate). As Ke-Hsueh Li pointed out (Li, 1994) (Li, 1992a), the concept of coherence is strictly linked to Heisenberg’s uncertainty principle, i.e., coherence space-time being actually equivalent to the uncertainty space-time. This is the range of space and time within which particles lose their classical features as individuality and countability (the operator become undefined). The particles and fields within coherence space-time range must be considered as an indivisible whole where phase is well-defined: thus, what occurs to “a part” of a CD, within its coherent space-time range, is occurring to the whole CD (Li, 1994). This is a noteworthy point also for overcoming the prevailing naïve picture of the HBs (Del Giudice, Galimberti, Gamberale, & Preparata, 1995) conceived still as forces among “particles”. As described, this classic idea originated from the 1st quantization can be fruitfully replaced by the QFT perspective (2nd quantization) where the apparent (non-directional) force is the emergent property deriving from an energy gradient which is NOT primarily tied to the bonding among molecules, but established is routed on the ground energy level (vacuum) (Preparata G. , 1995) as a consequence of the “em-field + matter-field” coupling over the whole high-numbered system, (Bono, Del Giudice, Gamberale, & Henry, 2012).
Another crucial topic is the one concerning ions and their solvation in water. Within an electrostatic conception of dissolution of electrolytes in water, the initial dynamics has no physical consistency, since few layers of water molecules should be able to keep some Na+ and Cl- ions apart from their crystal lattice when the energy barrier to be overcome in order to brake ion bonds is in the order of 5 eV and a single water layer could produce at most a dielectric drop of the Coulomb force equal to 13 ( εr = 13 and not εr = 80 which holds for the bulk). Again, only by abandoning an ingenuous “stics’n’balls” interpretation of condensed matter, and by taking into account the quantum electrodynamic nature of objects like ions and their coupling with vacuum, it is possible to describe consistently the spontaneous process of solvation showing that ions establish in the incoherent fraction of water their own coherence domains, with their energy gaps (bigger than the ion-bond energy), dissolving in the liquid phase of the solvent without collisions (Del Giudice & Preparata, 2000). This explains (i) why by increasing temperature the solvent power of water increases (despite the net value of bulk dielectric permittivity decreases), (ii) why there is no emission of bremsstrahlung radiation from an electrolyte solution and (iii) why the phenomenon of ion-cyclotron resonance occurs (Del Giudice, Fleischmann, Preparata, & Talpo, 2002).
There are numerous other cases, which we will only briefly mention here, as they go beyond the scope of this topic and will therefore be dealt with in future papers. These regard the morphogenic role of water in biological matter (Henry M. , 2020), interfacial water (Pollack, 2013), dispersion properties of biologically bound water upon exposure in the 10 Hz to 100 GHz range (Schwan, 1977), burning salt water upon RF-exposure (Roy, Rao, & Kanzius, 2008), branching chain reaction of water (Voeikov, 2010), coherent water and cellular information processing (Henry M. , 2015) as well as stable water mixtures of both hydrophobic/hydrophilic liquids (Germany Patent No. DE 1.557.213, 1966) by Viktor Schauberger.
-
3According to the slowly varying envelope (SVE) approximation the frequency spectrum of the “envelope amplitudes” of the em-field is concentrated only on one mode, |ω| ≪ ωk = |k| (in natural units). Doing so means to neglect the third order time-derivative term in the equations of motion, which shows an instability of the perturbative ground state (PGS) in the matter-em field coupling, and is responsible for a departure from it towards a non-trivial solution of the equation of motion: a coherent state (see (Bono, Del Giudice, Gamberale, & Henry, 2012) for further details).