Very Dilute Aqueous Solutions — Structural and Electromagnetic Phenomena


Yinnon TA1*

1K. Kalia, D.N. Kikar Jordan 90666, Israel

*Correspondence E-mail: lwcdsrc@kalia.org.il

Keywords: Water molecule aggregates; serially diluted aqueous solutions; electromagnetic fields; domains; quantum electrodynamic domains; coherence domains; ferroelectric orderings.

Received: November 23, 2016; Revised: September 6, 2017; Accepted: September 15, 2017; Published: November 24, 2017; Available Online: November 24, 2017.

DOI: 10.14294/WATER.2017.4

 

Abstract

Very dilute aqueous solutions have properties which are of importance for wastewater treatment, toxicology and pharmacology. Such solutions typically have concentrations below about 10-6 – 10-10 mol/liter. According to the customary theories, such solutions and distilled water should have similar structural, physicochemical and biological properties. However, previous experimental studies revealed that this is not always the case. When these solutions are prepared by serial dilutions and vigorous shaking after each dilution step, their properties may considerably differ from those of distilled water. Their properties may also significantly differ from those of serially diluted solutions with the same chemical composition, which were not vigorously shaken after each dilution step. In the current study, the following phenomena of serially diluted, vigorously shaken aqueous solutions are analyzed: (a) their self-similar topology, (b) their emission of ultra low frequency (ULF) radiation; (c) 7.85 Hz alternating magnetic fields affecting their structure, and (d) their structure after dilution below 10-24 mol/liter. Since ambient electromagnetic fields affect these phenomena, the analyses are carried out within the context of quantum electro-dynamics (QED). The analyses show that the QED model for serially diluted, vigorously shaken aqueous solutions, as developed by Yinnon and Yinnon [Int J Mod Phys B 25:3707-3743 (2011)], provides consistent explanations for phenomena (a)-(d).

 

Introduction

During the last two decades, research on serially diluted, vigorously shaken aqueous solutions intensified, as reviewed by Konovalov and Ryzhkina (2014) and Elia et al. (2015). These solutions are prepared from stock solutions with concentration (C) in the ~4 – 10-3 mol/liter (M) range. The stock solutions are repetitively (serially) decimally or centesimally diluted, at ambient conditions. After each dilution step, the liquids are vigorously shaken with a lab dancer shaker, by vertical vortexing or other methods.

The study of serially diluted, vigorously shaken aqueous solutions was stimulated by their now amply demonstrated significance for water reclamation, toxicology and medicine. The bioactivity of solutions of many (but not all) inorganic-, organic- or bio-molecules non-monotonically changes upon serial diluting to low (picomolar) or ultralow (femtomolar) concentrationsor beyond (Burlakova et al., 1986, 2004, 2005; Davenas et al, 1988; Palmina et al., 1994; Pynzar et al. 1995; Lobyshev et al., 2005; Ryzhkina et al., 2010, 2011a; Belov et al., 2011; Konovalov, 2013; Mishina et al., 2015; Voeikov and Yablonskaya, 2015). The effects of serially diluted, vigorously shaken aqueous solutions on bio-systems may follow an upside-down U-shaped concentration dependent response curve. Such a curve is observed for many but not for all solutes. The slope reversal appears at a solute type dependent threshold concentration (Cthr). Typically, Cthr is of the order of about 10-6 – 10-10 M.

Compounds released by glass containers are present in serially diluted, vigorously shaken aqueous solutions (Witt et al., 2006). Such compounds and other contaminants affect the physicochemical variables of these liquids (Elia and Niccoli, 2004a&b). However, in depth analyses of such liquids prepared in glass or plastic vessels show that such compounds and contaminants alone cannot account for their typical properties (Elia and Niccoli, 2004a&b; Ciavatta et al., 2008; Montagnier et al., 2009; Elia and Napoli, 2010; Upadhyay and Nayak, 2011; Demangeat, 2009, 2010; Pershin et al. 2015; Yinnon and Liu, 2015a,b).

Many aspects of serially diluted, vigorously shaken aqueous solutions have been measured by independent research groups. The main aspects are cited in the next section. The most pronounced aspect is that for some (but not for all) solutes, at C below Cthr, their physicochemical, structural and bioactive properties differ from those of serially diluted solutions with equivalent chemical composition, which were not vigorously shaken after each dilution step (Elia and Niccoli, 1999, 2000, 2004a; Elia et al., 2004b). Solute attributes underlying the differences are not yet clarified (Konovalov, 2013; Ryzhkina et al., 2015a; Yinnon and Liu, 2015c). In this paper, the denotation SDVSAS (serially diluted, vigorously shaken aqueous solutions) will be used for referring to the following liquids: Solutions which were prepared by serial dilutions and vigorous shaking after each dilutions step, for which at C<Cthr, their measured physicochemical properties significantly differ from those of serially diluted solutions with equivalent chemical composition which were not vigorously shaken after each dilution step.

Modeling SDVSAS requires a quantum electro-dynamic (QED) approach. Electromagnetic (EM) radiation majorly affects the structure and physicochemical properties of these liquids (Ryzhkina et al. 2011b, 2012a-c; Konovalov, 2013). For example, for C<Cthr, ambient EM radiation underlies formation of their 10-7 m sized nano-associates. Such large molecular associates are manifestations of QED effects (Preparata, 1995; Yinnon and Liu, 2015a; Fiorini, 2016). The QED model of SDVSAS, proposed by Yinnon and Yinnon (2011), has provided consistent explanations for many of their phenomena (Yinnon and Elia, 2013; Yinnon and Liu, 2015b,c).

The goal of this paper is to employ the QED model for clarifying some hitherto unexplained characteristics of SDVSAS reported in the literature. In-depth explanations for ultra violet (UV) and infra red (IR) spectral features of SDVSAS will be presented in a separate publication. Some reported characteristics will not be analyzed. Seemingly not all researchers are familiar with the hysteretic and far-from-equilibrium dissipative system properties of SDVSAS. These properties have implications for assuring SDVSAS phenomena are significant. For example, the following parameters affect the  physicochemical properties of SDVSAS samples: volume, storing conditions, the passage of time after preparation of the samples, shaking procedure, ambient EM radiation, temperature, chemical composition of the stock solutions from which SDVSAS are prepared and the purity of the water used for its dilutions (Elia et al., 2004a, 2008a&b, 2012; Lobyshev et al., 2005; Yinnon and Elia, 2013; Rhyzkina et al., 2010, 2011a-c, 2012a-c, 2013, 2015c; Konovalov and Ryzhkina, 2014; Mishina et al., 2015). The group of Miranda et al. (2011) demonstrated that the electrical impedance of SDVSAS of LiCl samples varies when prepared with distilled water procured from different laboratories. The group carefully controlled the SDVSAS preparation procedures, their storing conditions and their impedance measurements. Another implication of the hysteretic and far-from-equilibrium dissipative system properties of SDVSAS is that the controls (blanks) should be carefully designed. Perturbing pure water with serial dilutions and vigorously shaking after each dilution step also alters their structure and physicochemical properties (Elia et al., 2000, 2004a; Lobyshev et al., 2005; Rhyzkina et al., 2015b).

The outline of the paper is as follows: It starts with a concise overview of the main measured properties of SDVSAS. The overview is important for appreciating the issues presented in the subsequent sections. Next, in the “Theory Section”, the SDVSAS model is concisely summarized.a In the subsequent section, i.e., the “Analyses of Reported Data Section,” explanations are presented for the following phenomena: the self-similar topology of SDVSAS, the emission of ultra low frequency (ULF) radiation by SDVSAS, 7.85 Hz alternating magnetic fields affecting the structure of SDVSAS, and ordering of H2O in SDVSAS diluted below the Avogadro limit. After that, in the “Discussion Section,” the strengths and the shortfalls of the current version of the SDVSAS model are discussed.

In this section, also hypotheses concerning the bioactive principle in SDVSAS are addressed. No new experimental results are presented in this paper. All the experimental results, which will be analyzed, were obtained in previous studies and reported in the literature. A list with abbreviations is presented at the end of this paper.

a Elaborate discussions on the various aspects of the model are published in Yinnon and Yinnon (2011), Yinnon and Elia (2013) and Yinnon and Liu (2015b&c). Since QED of aqueous solutions has hitherto mainly been employed for explaining special phenomena, many readers may be unfamiliar with it. Its aspects relevant to this paper’s analyses are concisely summarized in Yinnon and Liu (2015a).

 

Measured Properties of SDVSAS

  1. At C<Cthr, molecular associates are present in SDVSAS (Elia and Niccoli, 1999, 2000, 2004a; Elia et al., 2004b). The associates are mainly composed of solvent molecules (Elia and Niccoli, 1999, 2000, 2004a; Elia et al., 2004b; Ryzhkina et al., 2011a-c, 2012a-c, 2013,2015a&b; Konovalov 2013; Konovalov and Ryzhkina, 2014). Conversely, in a serially diluted solution made of the same solute but which was not vigorously shaken, at C<Cthr, all their molecules are randomly distributed and participate in the Brownian movement.

  2. Hysteretic properties (Samal and Geckeler, 2001; Elia et al., 2004a, 2005, 2008 a&b; Lobyshev et al., 2005; Belon et al., 2008; Konovalov and Ryzhkina, 2014; Miranda et al., 2011).

  3. For C<Cthr, the electrical conductivity (χ) and the heat of mixing with acids or bases (Q) are linearly correlated (Elia et al., 2004b; Belon et al., 2008; Elia et al., 2015). At this C range, the pH and log χ are also linearly correlated (Elia et al., 2015). The correlations indicate that a single cause underlies the specific properties of SDVSAS (Elia et al., 2015, 2016).

  4. The quasi-periodic variations with macroscopic times (months and years) of the χ, Q and pH values of SDVSAS indicate that these liquids are far-from-equilibrium dissipative systems (Elia et al., 2000, 2008a&b; Belon et al., 2008; Elia et al., 2015). Hence, SDVSAS properties are not quantitatively reproducible but are qualitatively repeatable. It should be stressed that SDVSAS properties depend on the time of their storage, the volume of the samples and ambient EM fields.

  5. For C<Cthr but above 10-24 M (i.e., above the so-called Avogadro limit), EM radiation crucially affects the structure, physicochemical properties and bioactivity of SDVSAS (Ryzhkina et al. 2011b, 2012a-c; Konovalov, 2013). These SDVSAS lose their characteristic properties when stored for many hours (24 h or more) in a Permalloy container, i.e., under hypo electro-magnetic conditions (Konovalov and Ryzhkina, 2014). Within Permalloy containers the residual field is about 10 – 50 nano Tesla. In other words, on shielding these SDVSAS from some ambient EM radiation, their properties become the same as those of solutions with equivalent chemical composition but which were not prepared by serial dilutions and vigorous shaking after each dilution step.

  6. The dependence of the bioactivity of SDVSAS on the number of dilution steps (NDS) is interrelated with their physicochemical and structural properties (Lobyshev et al., 2005; for review see Konovalov and Ryzhkina, 2014).

  7. The sizes of the associates in SDVSAS at C<Cthr are in the range of 10-7 – 10-4 m (Ryzhkina et al., 2010, 2011a-c, 2012ac-c, 2013; Konovalov and Ryzhkina, 2014; Elia et al., 2015). The associates underlie the peculiar properties of these SDVSAS. Second order phase transitions typify their formation (Pershin et al., 2016). Since SDVSAS are far-from-equilibrium dissipative systems, the associates constantly reform. Perturbations (e.g., dilutions and shaking) reorganize their molecules. The reorganization is reflected foremost in the non-monotonic dependence of the dielectric permittivity of SDVSAS on C (i.e., this variable often is significantly affected by each dilution step and by the shaking of the liquid (Ryzhkina et al., 2011c, 2012b, 2013; Lunev et al., 2014; Yinnon and Liu, 2015c)). The reorganization also underlies the non-monotonic dependence of other physicochemical variables of SDVSAS on NDS [e.g., the surface tension, χ, the size and electrokinetic potential (ζ-potential) of the associates (Konovalov and Ryzhkina, 2014; Yinnon and Liu, 2015c)]. In addition, the reorganization underlies the non-monotonic dependence of the bioactivity of SDVSAS on NDS. The aforementioned non-monotonic dependencies are interrelated (Konovalov and Ryzhkina, 2014). The various non-monotonic dependencies have been observed by several independent research groupsb and were shown not to be attributable to contaminants.

 

b Miranda et al. (2011) showed that the electrical impedance of SDVSAS of lithium chloride (LiCl) depends non-monotonically on these liquids’ NDS. Duval et al. (2012) showed that the Landau-Placzek ratio, i.e., the ratio of the Rayleigh scattering intensity over the total Brillouin intensity, for SDVSAS of LiCl and for SDVSAS of sodium chloride (NaCl) depends non-monotonically on these liquids’ NDS. Pershin et al. (2015, 2016) demonstrated that for SDVSAS of potassium phenosan, the following variables were all found to depend non-monotonically on this liquid’s NDS: the frequency of the center of the Raman scattering OH vibration band of its H2O, the Rayleigh line width in its four-wave mixing spectra, its integrated elastic scattering profile of 527-nm laser pulses, its fluctuation of the elastic-scattering coefficient and its optical activity. Kozumi  and  Kitagawa (2016) demonstrated that the aquaporin permeability of serially diluted, vigorously shaken water kept for 24 h in contact with ceramics non-monotonically changes with the NDS. This liquid was 15 times centesimally diluted. After each dilution step, the aquaporin permeability changed non-monotonically.

 

Theory

The customary models of water and their solutions predict that serial dilutions, combined with vigorous shaking and exposure to ambient EM radiation, do not affect their characteristics (Horne, 1972; Robinson and Stokes, 2002). In addition, these models predict that H2O molecules move randomly, with the exception of solvation shells’ H2O. Moreover, these models predict that solvated solutes distribute homogenously, move independently and randomly. These customary models explicitly describe electro-static forces and assume that electro-dynamic ones can be treated perturbatively or often even may be ignored. However, QED models explicitly (non-perturbatively) describing electro-dynamic forces show that these forces may lead to association of H2O and solute molecules (Del Giudice et al., 1988, 1998, 2000, 2002, 2006; Preparata, 1995 chapters 2, 5, 10; Arani et al., 1995; Bono et al., 2012; Yinnon and Yinnon, 2012). Interactions between EM radiation and electrolytic solutes, polar solute molecules, the dipole moments of H2O or the electrons of H2O may lead to formation of various domain types. Here such domains are denoted QED domains. Formation of these domains occurs only in specific concentration ranges, which depend on solute type. These domains may agglomerate into supra-domains. These supra-domains are not ensembles of molecules but agglomerates of domains, like domains in liquid crystals.

A concise review of the QED theory of water and their solutions has been presented by Yinnon and Liu (2015a). The various hitherto identified QED domains present in these liquids, their schematic pictures, their properties and the physics underlying the dependence of the domains’ formation on concentration are all summarized in Yinnon and Liu (2015a). In their review, Yinnon and Liu (2015a) emphasized that the QED theory of water is based on an explicit non-perturbative description of the van der Waals’ dispersion forces. The imperative for adequately describing van der Waals’ dispersion interactions and other electro-dynamic interactions in various condensed matter systems has also recently been pointed out by Ferri et al. (2015), Fiorini (2016) and Ambrosetti et al. (2016).

 

Figure 1. This figure presents a schematic picture of serially diluted solutions of weak electrolytes or non-electrolytic compounds. The top row (a) and bottom row (b) series pertain to solutions which, respectively, were not vigorously shaken and those which were vigorously shaken after each dilutions step. Figure ia illustrates that for C larger than a transition concentration , all solvated solutes move randomly, i.e., do not organize in a QED domain. The tiny blue balls represent randomly moving ~10-9 – 10-8 m solvated solutes. The irregularly shaped bunches of black hexagons represent aggregates of non-solvated solutes. Figure iia illustrates that on dilution below , solvated solutes organize in a QED domain type denoted CDplasma (symbolized with purple-blue colored balls). The yellow-brown balls and their agglomerates represent, respectively, ~10-7m sized QED domains denoted and  supra-, which both are stabilized by CDplasma. Figures iia and iiia illustrate the transformation of CDplasma into another type of QED domain denoted IPDplasma. The transformation occurs at the transition concentration . Figures ia-iiia illustrate that on dilution the non-solvated solutes diminish, i.e., solvate. Figures iiia-va illustrate that on dilution the diameter of IPDplasma does not change, but the number of IPDplasma diminishes. Figures via-viiia illustrate that below a certain concentration there are insufficient solutes to form IPDplasma. The concentration below which no IPDplasma form has yet not been theoretically derived. Figures via-viiia illustrate that whenever there are too few solutes to form IPDplasma the solution has the characteristics predicted by the customary models, i.e., all solvated solutes move randomly and their number diminishes on dilution.

In the Figure 1b series, the blue zigzag curves symbolize that shaking excites or cracks domains and aggregates. Figure iib illustrates that excitations or cracking does not significantly alter the internal structure of CDplasma, which just as in Figure 1iia are represented with purple-blue colored balls. Figures iib and iiib illustrate the transition from CDplasma to IPDplasma, with the latter pictured as blue-crystalline balls (just as in Figure iiia). Figures iiib and ivb illustrate that shaking excites or breaks up IPDplasma. The excited or broken IPDplasma pieces, which in the text are denoted electric dipole aggregates (EDAIPDplasma),are pictured as irregularly shaped aggregates in figure ivb. Their aligned black arrows orderings symbolize EDAIPDplasma’s distorted ferroelectric H2O orderings. The purple arrow in the EDAIPDplasma symbolizes these domains’ dipole moments. Figures ivb and vb illustrate that on diluting below a solute type dependent critical concentration () a QED domain denoted CDrot becomes stabilized by EDAIPDplasma, i.e., the irregular shaped EDAIPDplasma are located within the elongated ovals representing CDrot. The dark blue arrows symbolize the dipole moment of CDrot. Figure vib shows that vigorous shaking excites or breaks up CDrot, thus creating entities denoted EDACDrot. The lump outlined with an irregular shaped broken curve and located at the bottom of one of the left CDrot, represents the EDACDrot. Figures vib-viib show that at certain concentrations both EDAIPDplasma and EDACDrot are present within CDrot,  though the sizes of EDAIPDplasma diminish with concentration. Figure viiib shows that on diluting further, no EDAIPDplasma persist, i.e., there are too few solute particles to sustain EDAIPDplasma. At these concentrations, vigorous shaking just breaks up CDrot and creates new EDACDrot. These in turn stabilize new CDrot, as pictured in Figure viiib. Figures vb-viib illustrate that CDrot may align with their dipole moments parallel. Figure viiib illustrates that at certain concentrations their dipoles may be aligned anti-parallel. (Note that the sizes of the various domains, their broken pieces and the sizes of the solvated solutes with their hydration shells are not presented according to their realistic scale ratios.)


 

SDVSAS MODEL

Fig.1 presents a schematic picture of the SDVSAS model proposed by Yinnon and Yinnon (2011) (i.e., the structure of SDVSAS for different C ranges). The figure is a reprint of Fig.1 in Yinnon and Liu (2015c). The figure pertains to SDVSAS of weak- or non-electrolytic compounds. The schematic picture for the model of SDVSAS of strong electrolytes is very similar. It appeared in Fig. 1 of Yinnon and Liu (2015b). The differences between the models of weak-, non- and strong electrolytes are minor ones, as discussed below in paragraphs I and II. The various aspects of the models are discussed in the following paragraphs I-VI and i-vii. Experimental results, obtained in independent research groups, agree with these aspects, as shown by Yinnon and Yinnon (2011), Yinnon and Elia (2013) and Yinnon and Liu (2015b&c). The few aspects which still have to be definitely confirmed by measurements will be explicitly pointed out below.

  1. In solutions of weak- or non-electrolytes, for C above a transition concentration , only few solute molecules solvate. The solvated solutes move randomly (see Fig.1i). For < C < , part of the solvated weak electrolytes, or solvated non-electrolytes with a sufficiently large electric dipole moment, aggregate in a QED domain type denoted CDplasma (see Fig.1 ii). Within these domains, the solute molecules together with their solvation shell H2O perform coherent plasma oscillations. The diameter of a CDplasma is of the order of 10-6 m. CDplasma may agglomerate into supra-CDplasma. A schematic picture of CDplasma is presented in Fig. 3a in Yinnon and Liu (2015a). QED theory predicts that the interactions between the solute molecules and Tera Hertz (Hz) to Mega Hz EM radiation underlie formation of CDplasma, but this has not yet been confirmed by experimental data. The transition concentrations and depend on solute type. Typically, ~10-4 M < < 1M and ~10-6 M < < ~104M.

    In solutions of strong electrolytes, at any C larger than , part of the solvated solutes together with numerous H2O organize in CDplasma. Thus, the difference between solutions of weak- or non-electrolytes versus those of strong electrolytes is that in the latter there exist no concentration above which CDplasma do not form, i.e., there is no .

    According to QED theory, CDplasma and supra-CDplasma may stabilize a QED domain type denoted . These domains may aggregate into supra-. A is composed of H2O, which coherently oscillate between their electronic ground state and an excited electronic state. Interactions between H2O and UV radiation underlie formation of . The diameter of a is about 10-7 m. These domains cannot contain solutes. A schematic picture of is presented in Fig. 1 in Yinnon and Liu (2015a).

    The abovementioned characteristics for C > hold independent of the solutions’ preparation procedure. In other words, these do not just hold for SDVSAS, but also for solutions prepared without serial dilutions or vigorous shaking. Serial dilutions or vigorous shaking affect CDplasma and , mainly causing their breakup. However, after perturbations are over these domains reform, as illustrated in Figs.1 iia&b.

    The energetics of processes induced by vigorous shaking of SDVSAS have been evaluated by Yinnon and Liu (2015b) on pages 59-60 (footnote g). The evaluation indicates that vigorous shaking provides about 1010 – 1015 eV per aggregate (domain). The energy required for desorption of a molecule from a is of the order of 10-1 eV, while a few eV are required for desorption of a molecule from a CDplasma. Thus, the energy applied to SDVSAS by vigorously shaking is sufficient to breakup and CDplasma.

  2. In solutions of strong electrolytes, at C , CDplasma transform into another type of QED domain, i.e., IPDplasma (see Figs.1ii-iii). QED theory predicts that the interactions between the solute molecules and Tera Hz to Mega Hz EM radiation underlie formation of IPDplasma, but this has not yet been confirmed by experimental data. IPDplasma, just as CDplasma, are composed of few solvated solutes and numerous H2O. The solvated solutes in IPDplasma are crystalline ordered. The solvated solutes and H2O constituting IPDplasma perform in-phase plasma oscillations. A schematic picture of IPDplasma is presented in Fig. 3b in Yinnon and Liu (2015a). The diameter of an IPDplasma is about 10-6 m. IPDplasma may aggregate in supra-IPDplasma. Dilution below diminishes the number of randomly moving solvated solutes as well as the number of solvated solutes incorporated in IPDplasma. At very low concentrations, the number of solutes is too low for formation of IPDplasma. As shown in Fig. 1iiib, IPDplasma may stabilize and supra-.

    The abovementioned characteristics for C < hold independent of the solutions’ preparation procedure, i.e., not just for SDVSAS, but also for solutions prepared without serial dilutions or vigorous shaking.

    For solutions of weak- or non-electrolytes, for C , QED theory also predicts that CDplasma transform into IPDplasma. The differences between the transition concentrations and may be tiny. However, definite experimental evidence for these aspects of such solutions has yet to be provided, as discussed by Yinnon and Liu (2015c).

  3. For solutions containing IPDplasma, their agitation by vigorous shaking affects their properties. Vigorous shaking transforms IPDplasma into aggregates with an electric dipole moment (see Fig.1iv). These aggregates are denoted electric dipole aggregates EDAIPDplasma. The sizes of EDAIPDplasma may reach 10-6 m, but may also be much smaller.

    The formation of EDAIPDplasma results from agitations exciting or breaking up IPDplasma. A few eV are required for desorption of a molecule from an IPDplasma. Since vigorous shaking provides about 1010 – 1015 eV per (domain) (Yinnon and Liu, 2015b), vigorous shaking of SDVSAS can break up IPDplasma.

    The excitations or break up of IPDplasma partly destroy the spherically symmetric alignments of the dipole moments of the numerous H2O in the large solvation shells surrounding its crystalline ordered solvated solutes. The partial destruction of the spherically symmetric alignments endows the disturbed IPDplasm with an electric dipole moment.

    Vigorous shaking of SDVSAS containing CDplasma, i.e., at C > , does not create electric dipole aggregates. Solutes are randomly distributed in a CDplasma, i.e., are not crystalline ordered. Moreover, only few H2O constitute its solvation shells, i.e., only few H2O are spherically symmetric aligned. Therefore, excitation or break up of CDplasma does not create aggregates with a sizable electric dipole moment.

  4. EDAIPDplasma may stabilize and supra- (see Fig.1 iv).

  5. For C less than the critical concentration , yet another type of QED domain may form, i.e., CDrot. This domain is composed of ferroelectric ordered H2O. A CDrot has an electric dipole moment due to the ferroelectric ordering of its H2O. The H2O constituting CDrot coherently oscillate between two of their rotational states. Interactions between the H2Os’ electric dipole moments and IR EM radiation underlie formation of CDrot. For a schematic picture of CDrot see Fig. 2 in Yinnon and Liu (2015a). In bulk water at ambient conditions, CDrot do not auto-organize. The energy gained by a H2O joining a CDrot is of the same order as the energy of the thermal fluctuations. Therefore, thermal aggression prevents auto-formation of CDrot at ambient conditions. However, objects with sizable asymmetric charge distributions (e.g., macromolecules or hydrophylic membranes) may induce formation of CDrot (i.e., stabilize CDrot). Their stabilization causes a permanent time-dependent polarization of the liquid. The sizes of CDrot may increase with the passage of time. At ambient conditions, their stabilization, such as by hydrophilic membranes, may require many hours or days. Their sizes may reach 10-4 m.

    Solutes are pulled into CDrot. Few solute particles can locate in CDrot and not wreck their host. Many solute molecules destroy CDrot. The critical concentration below which CDrot persist () depends on the solute, typically ~10-10 M < < ~10-6 M. Analyses of experimental data of many aqueous solutions have shown that  equals Cthr.

    For C < , due to the interactions between the electric dipoles of EDAIPDplasma and the electric dipoles of H2O, EDAIPDplasma stabilize CDrot (see Fig.1vb). In other words, EDAIPDplasma, due to their significant asymmetric charge distributions, stabilize CDrot. Stabilization of CDrot also may be induced by solutes with sufficiently large permanent or induced electric dipoles. Even when no IPDplasma and hence EDAIPDplasma form in SDVSAS, solvated solutes with a sufficiently large asymmetric charge distribution still might induce CDrot stabilization for C < .

    Vigorous shaking excites or breaks up CDrot (see Fig.1v). The energy required for desorption of a molecule from a CDrot is of the order of 10-2 eV.  Since vigorous shaking provides about 1010 – 1015 eV per (domain) (Yinnon and Liu, 2015b), vigorously shaking of SDVSAS can break up CDrot. Due to the ferroelectric ordering of the H2O constituting CDrot, excited or broken CDrot also are electric dipole aggregates, which are denoted as EDACDrot (see Fig.1vi). Due to interactions between the electric dipoles of EDACDrot and the electric dipoles of H2O, EDACDrot also stabilize CDrot. Therefore, serial dilutions with vigorous shaking at each dilution step diminish solvated solutes and EDAIPDplasma, but EDACDrot persist. These EDACDrot stabilize CDrot and supra-CDrot too. As a result, CDrot persists up to ultra-low concentrations and beyond (see Figs.1vib-viiib).

    CDrot and supra-CDrot may stabilize and supra-. Such agglomerates are denoted [supraCDrot <supra->].

    CDrot may agglomerate into supra-CDrot with their dipole moments more or less parallel or anti-parallel oriented. The vigorous shaking applied to SDVSAS after each dilution step restructures the orientations of the dipole moments of its supra-CDrot. The restructuring strongly affects the properties of SDVSAS. For instance, it underlies the non-monotonic dependence of its physicochemical variables on C. Moreover, it affects the stabilization of  and the formation of supra-.

    According to the QED model of SDVSAS, when these liquids are slightly or extremely diluted below , the entities underlying their extraordinary physicochemical properties are CDrot, EDACDrot and . In addition, it has been hypothesized that for C = Cthr, underlies the bioactivity of SDVSAS. Therefore, on screening SDVSAS from the ambient EM radiation that mediates formation of their CDrot and , these domains disintegrate and the SDVSAS lose their typical characteristics. Placing SDVSAS in Permalloy containers screens the EM radiation which mediates formation of their CDrot and .

Explaining some measured properties of SDVSAS requires knowledge of several characteristics of . Therefore, in the following paragraphs, these characteristics will be concisely summarized from the papers published by Preparata (1995), Arani et al. (1995), Del Giudice and Preparata (1998), Del Giudice et al. (2002, 2010a&b), Del Giudice (2007), Bono et al. (2012) and Montagnier et al. (2011, 2015).

  1. Within a , the photons underlying the electronic transitions of its H2O are condensed (Bose-Einstein condensation). These photons mediate the attractions between its H2O. These condensed photons are trapped and cannot be irradiated outward. Or more accurately, only when a  disintegrates, for example due to heating, the trapped photons “evaporate.”

  2. The H2O constituting a  coherently oscillate between their ground electronic state and an excited electronic state |b>. The energy H2O gain by joining a  [i.e., the energy (Ecoh) a H2O gains by participating in the coherent oscillations] depends on temperature. At ambient conditions, Ecoh is of the order of 10-1 eV.

  3. For bulk water at ambient conditions, the energy of its thermal fluctuations is about an order of magnitude smaller than Ecoh. Therefore, its  are meta-stable. The competition between QED interactions and collisions (thermal noise) causes a continuing adsorption and desorption of H2O from its . In other words, there are everlasting transitions of H2O between a coherent and a non-coherent phase. The transitions are like those in superfluidic Helium. The transitions produce a continuously changing space distribution of the coherent and non-coherent fractions of H2O. For bulk water at ambient conditions, only about 20 percent of its H2O belong to the former; the time scale of the transitions is below 10-14 seconds (s). Therefore, observation of a  requires very fast probes. Indirect experimental evidence for presence of  in bulk water has been provided by De Ninno and Congiu Castellano (2011). Experimental data indicating stabilization of  by membranes or other QED domain types has recently been forwarded (Elia et al., 2015, 2017; Yinnon et al., 2016; Yinnon and Liu, 2015b,c).

  4. Whenever external forces affect a and the energy transferred to the domain is less than Ecoh, due to the coherence, the whole domain excites. In other words, these forces do not just perturb one of its molecules but excite the whole domain.

  5. contains many coherent quasi-free electrons (QFE), because its H2O molecules oscillate between the ground electronic and the nearly free |b> electronic states.c Exciting a  with energies less than Ecoh , e.g., 0.17 eV at ambient conditions (4×1013 Hz), leads to coherent vortices in its pool of QFE. Due to their coherence, the vortices have no internal friction. The vortices cannot decay thermally because the coherence prevents random (thermal) fluctuations. In other words, the vortices are cold. Therefore, their lifetimes are very long and depend on those of the , e.g., the vortices may follow the dissipative dynamics of the .

    The energy spectrum of the vortices of the QFE of  is typified by a lowest lying excited state with a rotational frequency of the order of 1000 Hz. The energy spacing of the spectrum is of the same order, i.e., in the ultra low frequency (ULF) range. The magnetic dipoles of these vortices may become aligned by external magnetic fields, e.g. the terrestrial magnetic field.

  6. The phases of the coherent oscillations of the H2O of different may become correlated by electric or magnetic fields, e.g., the field due to the electric dipole moment of a CDrot or external fields. Such correlation leads to a coherence among , which enhances Ecoh. Therefore, the presence of such fields may stabilize , e.g., on placing bulk water in external fields may enhance the prevalence of its  .

  7. Ions close to a  get attracted by the fields of the photons condensed in the domain. When the motion of these ions does not get altered by collisions with other molecules, they orbit around the . The motion of ions incorporated in CDrot, CDplasma or IPDplasma does not get altered by collisions. The energy such ions gain by their participation in the coherent dynamics of these domains protects them from the thermal aggression of neighboring molecules. Hence, when a  is stabilized by a CDrot, CDplasma or IPDplasma, part of the ions present in these domains orbit around .

    The circular speed of ions orbiting around a  is proportional to the ion cyclotron frequency (νc). For νc holds νc=(qB)/2πm, where q and m are the electric charge and the mass of the ion, respectively, and B is the strength of the static magnetic field (e.g. the terrestrial magnetic field).

    On applying an alternating magnetic field with a frequency matching νc, ions are extracted from their orbits around . The extraction, due to the conservation of angular momentum, causes rotational excitation of the QFE of the , that is, formation of coherent vortices. When the ion concentration is about uniform in a mesoscopic region and an external field also spans over mesoscopic distances, the excitation energy of the vortices of many  may be uniform. A uniform excitation of leads to coherence among them, which in turn enhances their Ecoh and their stabilization. Recent experimental findings by D’Emilia et al. (2015, 2016) seemingly support this characteristic. The findings suggest that on exposing pure water to sufficiently strong magnetic fields with a frequency equaling that of one of the ion cyclotron frequencies of H3O+, e.g., νc=7.83 Hz, enhances stabilization of .

c Fingerprints of these quasi-free electrons endowing  with a negative charge were revealed by electrophoresis carried out by Konovalov and Ryzhkina (2014), as discussed by Yinnon and Liu (2015c).

 

Analyses of Published Experimental Data


Self Similar Topology of SDVSAS

Investigations of the dielectric permittivity and electrical conductivity of SDVSAS of membranotropic amphiphilic calix[4] resorcinarene with tris (hydroxymethyl) methylamide groups (MAC4RWTHMG) at concentrations in the range of 10-11-10-3 M have shown that its molecular associates have fractal structures (Lunev et al., 2014). Analyses by Yinnon and Liu (2015c) have shown that this SDVSAS has the characteristics specified in paragraphs I-V — the analyses pertain to the experimental data on the physicochemical properties of these liquids and on the sizes and electrokinetic potential of their molecular associates measured by Ryzhkina et al. (2012b).

Self similarity is a property of the QED domains (Vitiello, 2012, 2014), e.g., of supra-CDrot, supra-, supra-CDplasma and supra-IPDplasma. According to the above presented QED model, for C < , SDVSAS contain one or more of these domain types. Therefore, the molecular associates in SDVSAS, which are mediated by EM radiation, indeed should have fractal orderings. To the best of my knowledge, the analyses of the experimental data of the SDVSAS of MAC4RWTHMG, carried out by Lunev et al. (2014), provide the first confirmation of the self similarity aspect of the SDVSAS model. It should be noted, however, that the self similarity of the H2O orderings in water perturbed with a Nafion membrane (iterative Nafionized water) or in water perturbed by iterative filtrations was recently proven by Capolupo et al. (2014). These researchers showed that the linear correlations between the logarithm of various physicochemical properties (e.g., χ, Q or density) and the pH of these perturbed waters [measured by the group of Elia et al. (2013a,b, 2014a,b, 2015)] reflect the self similarity of their H2O orderings. These orderings have been identified as and CDrot by Yinnon and Elia (2013), Yinnon et al. (2016) and Elia et al. (2017). Linear correlations between the logarithm of χ, the logarithm of Q and pH have also been observed for SDVSAS (Elia et al., 2004b; Belon et al., 2008; Elia et al., 2015). Thus, from the analyses of Capolupo et al. (2014), Elia et al. (2004b), Belon et al.(2008) and Elia et al. (2015), one could infer that the H2O orderings observed in SDVSAS should have a fractal structure. The study by Lunev et al. (2014) confirmed this inference.

Emission of ULF radiation by SDVSAS

Data on SDVSAS emitting ULF radiation reported in the literature — Montagnier et al. (2009, 2011, 2015) demonstrated that SDVSAS of some types of deoxyribonucleic acid (DNA) emit ULF (500-3000 Hz) EM radiation.d The SDVSAS were prepared in plastic 1.5 mL Eppendorf tubes. Specifically, they demonstrated the following:

  • The number of dilution steps required for the SDVSAS to emit the radiation depends on the type of the DNA (e.g., DNA extracted from bacteria or viruses). Typically, at least six decimal dilutions of a few nano gram (ng) of DNA are required for the emission of ULF radiation to be detectable. After applying more than about 10 to 17 decimal dilutions, the emission is no longer observable. Expressed differently, the emission of ULF radiation is typically observed for SDVSAS of DNA with concentrations in the 10-5 M to 10-18 M range. The intensity of the emitted radiation non-monotonically depends on the concentration. The fact that no emission was detectable for SDVSAS of DNA with C<10-18 M might be due to limitations of the experimental set up. The emission might be too weak to be observable with the simple apparatus employed by Montagnier et al. (2009, 2011, 2015).d No emission of ULF radiation was detectable for serially diluted DNA solutions which were not vigorously shaken after each dilution step.
  • The SDVSAS of DNA only emit ULF radiation when these are stimulated by background extremely low frequency (ELF) radiation. The sources of the ELF radiation may be the natural Schumann resonances of the geomagnetic field or artificial ones. The frequency of the Schumann resonance, which has the lowest frequency, is 7.83 Hz (Nickolaenko and Hayakawa, 2002). Storing the SDVSAS in Permalloy containers, which screens the background EM radiation, affects their emission of the ULF radiation. For some samples, after several hours of such storage, no emission was detectable anymore. For other samples, after having been stored for 48 h under equivalent conditions, emission of ULF was still detectable.
  • The radiation emitted by the SDVSAS of DNA is absorbable by neighboring pure water samples. The absorbed radiation alters the physicochemical properties of the neighboring pure water samples. These samples acquire properties reminiscent of those of the SDVSAS of DNA.

 

Elia et al. (2012) demonstrated that also SDVSAS of other solutes (e.g., fullerene), which were 5, 7, 9, 12 or 30 times centesimally diluted, emit radiation which alter the physicochemical properties of neighboring water samples.

Explanations for the emission of ULF radiation by SDVSAS — Montagnier et al. (2011, 2015) provided explanations for some of their experimental results. Their explanations, which are based on the processes detailed in paragraphs (v) and (vii), are as follows.

a) DNA are poly-electrolytes surrounded by a cloud of positive counter-ions. These ions may orbit around  (Del Giudice et al., 2002). The circular speed of the orbiting ions is proportional to the ion cyclotron frequency (vc). Typically, vc falls in the 1-100 Hz range. Ions orbiting around  get extracted from their loops by background alternating magnetic fields (Schumann resonances or artificial fields). Extraction of the ions from their orbits (because of conservation of angular momentum) causes a rotation of the plasma of the quasi-free electrons (QFE) present in the  (i.e., the extraction excites the domain and induces vortices). Since the mass of electrons is much smaller than that of the ions, the rotating plasma has a frequency vp that is much higher than vc. Typically, the lowest vp value is of the order of 1000 Hz and the spacing between the rotating plasma frequencies is of the same order, e.g., in the ULF range.

b) The rotation of the QFE produces the magnetic component of the ULF radiation.

c) The frequency and intensity of the emitted ULF radiation depend on the concentration of the counter-ions, i.e., the degree of dilution of the SDVSAS of DNA. The following causes underlie this dependency: Firstly, vp depends on the number of counter-ions participating in the orbiting and extraction processes. This number is a function of their concentration. vp is also determined by the quantization of angular momentum. Secondly, the co-resonating fields appearing in the surroundings of the rotating depend on the concentration of the counter-ions. These fields may produce coherence among the individual and thus affect the frequency and intensity of the emitted ULF radiation.

Montagnier et al. (2011, 2015) emphasized that their conjecture — that the emission of ULF radiation is due to magnetic activity — is supported by the absence of such emission after the SDVSAS samples were stored in Permalloy containers.e

As to the solute reminiscent alterations of pure water samples neighboring on SDVSAS of DNA, Montagnier et al. (2011, 2015) conjectured that it is due to the SDVSAS samples emitting ULF radiation which triggers coherence among the of the pure water samples. They noted that the physics underlying the copying of properties of SDVSAS to neighboring pure water samples is  similar to that of: “the proximity effect observed in two superconducting samples or in the arrays of Josephson junctions, by which the samples or the junctions fall into a phase-locking regime.”

e EM radiation at frequencies of kHz – Tera Hz can also be attributed to vibrations of electrically polar structures (Cifra, 2015), e.g., proteins, EDACDrot or EDAIPDplasma.

Unexplained phenomena pertaining to emission of ULF radiation by SDVSAS — To the best of my knowledge, no explanations have yet been forwarded for several of the observations made by Montagnier et al.’s (2009) and Elia et al. (2012).  In the following, I attempt to clarify them with the QED model of SDVSAS:

♦ Why was emission of ULF radiation not detectable for serially diluted aqueous solutions of DNA, when these solutions were not vigorously shaken after each dilution step?

According to the QED model’s aspects summarized in paragraph I: For < C < , part of the solvated solutes aggregates in CDplasma. Organization of DNA molecules (or their counter-ions) in domains with characteristics of CDplasma has been observed by Samal and Geckeler (2001), as discussed by Yinnon and Yinnon (2009).

According to paragraph II: for C , CDplasma transform into IPDplasmaTypically 10-6 M < <10-4 M. According to paragraph III, vigorous shaking of solutions containing IPDplasma excites or break up these domains and creates EDAIPDplasma.

According to paragraph V: Firstly, for C , EDAIPDplasma stabilize CDrot. Secondly, vigorous shaking of the liquid excites or breaks up the CDrot and thus creates EDACDrot. Thirdly, the CDrot and EDACDrot may agglomerate into supra-CDrot with their electric dipole moments ordered more or less parallel or anti-parallel. The orientation of the electric dipole moments of the CDrot affects the physicochemical properties of the SDVSAS.

As noted in paragraphs I-V, may get stabilized by all the domains mentioned in the last three paragraphs, i.e., by CDplasma, IPDplasma, EDAIPDplasma, EDACDrot or CDrot. For C above the Avogadro limit, all these domains contain coherently oscillating ions, which can orbit around . Vigorous shaking may excite the , i.e., its plasma of QFE.

In regards to the information presented in the last four paragraphs and in paragraphs (a) – (c), the observation that only serially diluted solutions of DNA which were vigorously shaken after each dilution step emitted ULF radiation, alludes to the following: Only the micron sized CDrot, EDACDrot, EDAIPDplasma or their agglomerates (creation of which requires serial dilutions and vigorous shaking after each dilutions step, and which all contain ferroelectric ordered H2O) can induce coherence among many of the 10-7 m sized . The coherence assures that the intensity of the ULF emitted by the excited plasma of QFE of the is strong enough to be detected with a solenoid.

The abovementioned shows that the observed emission of ULF radiation by SDVSAS provides additional support for the main aspect of its model.  In other words, CDrot, EDACDrot, and EDAIPDplasma are the entities underlying these liquids’ extraordinary properties. It also substantiates the conjecture of Montagnier et al. (2015) that the emission of ULF radiation by SDVSAS of DNA is the outcome of symmetry breaking. As they formulate it: “the symmetry which gets broken is the rotational symmetry of the electrical dipoles of the H2O and correlation modes are the ones associated to the dipole waves (similar to spin waves in ferromagnets).”

Which ions orbit around  in SDVSAS diluted below the Avogadro limit?

Elia et al. (2012) demonstrated that SDVSAS diluted below the Avogadro limit emit EM radiation. Their experiments, however, did not reveal the wavelengths of the radiation. In case the radiation includes ULF radiation, the question arises: Which ions orbit around in these extremely diluted liquids? Elia et al. (2012) prepared these SDVSAS in glass vessels. Therefore, these liquids contained dissolved components of the glassware containers at impurity level concentrations. However, in the absence of such impurities, such as when SDVSAS are prepared in plastic containers, H3O+ or OH ions are present. The pH of SDVSAS is affected by their molecular associates (Elia et al., 2004a; Ryzhkina et al., 2010, 2012b, 2013). [The effect of the associates on the pH has been explained in paragraph vi of the discussion section in Yinnon and Liu (2015c).] Hence, H3O+, OH or impurity ions may all orbit around in SDVSAS.

Why is emission of ULF radiation no longer detectable for some samples, after their storage for a few hours in Permalloy containers, while other samples stored for 48 hours under equivalent conditions still emitted ULF radiation?

For clarification of these phenomena, it should be emphasized that experiments have shown that the dynamics of SDVSAS are that of far-out-of equilibrium dissipative systems (Elia et al., 2008a,b, 2010). The dissipative dynamics of their molecular associates occur over macroscopic time scales, i.e., the dynamics may continue for months and even many years when the SDVSAS are kept at room temperatures and pressures.

The dissipative dynamics of SDVSAS, according to the QED model of SDVSAS, are due to a constant reorganization of the CDrot and EDACDrot constituting the supra-CDrot (Yinnon and Elia, 2013). It is also due to the stabilization and disintegration of and reorganization of supra-. Agitating SDVSAS by vigorously shaking accelerates reorganization of its domains. It also enhances excitation of the QFE of its . The dynamics of CDrot (which is related to the rotational excitation of H2O) is much slower than that of (which is related to the electronic excitation of H2O).

The continuous reorganization of the domains in SDVSAS affects the radiation emitted and absorbed by these liquids. On disintegration of , the EM radiation condensed within these domains is released (see paragraph i). The released radiation may enhance stabilization of new ,  it may escape to the environment or it may transform to heat. The same holds for the EM radiation condensed within CDrot. Reorganization of the domains in SDVSAS also affects the vortices of the QFE of , as it affects the emission of ULF radiation. Part of the emitted ULF radiation may get absorbed by QFE of neighboring .

Placing SDVSAS in Permalloy containers screens its CDrot and  from ambient IR and UV radiation. These are the sources of the radiation mediating the interaction between the molecules constituting these domains (see paragraphs I and V). However, the EM radiation released by the reorganization and disintegration of CDrot and  cannot escape through the walls of the Permalloy container to the environment. Thus, on placing SDVSAS samples in Permalloy containers, the non-linear dissipative processes taking place in these liquids imply that in some samples the CDrot and persist for a long time and continue to emit ULF radiation.  Conversely, in other samples most of these domains quickly disintegrate.

Why does the intensity of the ULF radiation emitted by the SDVSAS of DNA depend non-monotonically on their concentration?

The non-monotonic dependence on concentration of the intensity of the ULF radiation emitted by SDVSAS [as with similar dependencies of other physicochemical properties of SDVSAS, noted in paragraph (7)] result from the reorganization of their associates. More specifically, according to the model of SDVSAS, every perturbation like a dilution or shaking alters the ordering of the electric dipole moments of the supra-CDrot. These alterations affect the prevalence of the  and supra- . As such, the alterations also affect the ULF radiation emitted by the pool of quasi free electrons of .

Hitherto, to the best of my knowledge, emission of ULF radiation has been reported only for SDVSAS of some types of DNA. This calls for research aimed at searching for SDVSAS of other solutes which emit ULF radiation. Such research might verify the various explanations presented in this section.

 

Impact of EM radiation on the formation of associates in SDVSAS

This section shows that recent experimental data, obtained by A.I. Konovalov’s group (D.A. Konovalov et al., 2015), support some aspects of the explanations of Montagnier et al. (2011, 2015) discussed in the previous section. The data supports the hypothesis that the  present in an aqueous SDVSAS with concentration less than Cthr but above the Avogadro limit are affected by 7.85 Hz EM fields.

A.I. Konovalov’s group was the first to demonstrate that only when SDVSAS with C less than Cthr but above the Avogadro limit were stored under ambient EM fields, the presence of molecular associates in these liquids were observable with dynamic light scattering (DLS) (Ryzhkina et al., 2011a; Konovalov and Ryzhkina, 2014). After storage of these SDVSAS for 24 hours in Permalloy containers, DLS could not distinguish associates in these liquids. Moreover, after such storage, these liquids’ characteristic physicochemical and bioactive properties disappeared. Analyses of the associates and their impact on the physicochemical properties of SDVSAS indicate that these associates include , CDrot and EDACDrot (Yinnon and Liu, 2015b&c).

Recently, A.I. Konovalov’s group built a hardware complex, which enabled controlling the amplitude and frequency of the magnetic field in a Permalloy container (D.A. Konovalov et al., 2015). The group prepared samples of SDVSAS of biologically active p-sulfonato-calix[6]arene with C=1×10-9 M. The Cthr of this SDVSAS is about 10-7 M, significantly larger than C. One series of the samples was kept for 24 hours on the laboratory bench. In this paper, these samples are denoted as SDVSASLB. A second series was kept for 24 hours under hypo electromagnetic conditions, in the Permalloy container. Here these samples are denoted as SDVSASP. A third series was placed for 24 hours in the hardware complex, with the frequency of its magnetic field tuned to 7.85 Hz. These samples are denoted as SDVSASMF-7.85. This third series was subdivided into four sub-series, each of which was subjected to magnetic fields with different amplitudes, respectively, 48, 24, 12 and 6.4 A/m. These samples are denoted as SDVSAS48, SDVSAS24, SDVSAS12 and SDVSASL6.4, respectively.

D.A. Konovalov et al. (2015) showed that molecular associates were detectable with DLS in their SDVSASLB, but were not detectable in their SDVSASP. The average hydrodynamic diameter of the associates present in the SDVSASLB samples was 230 nm. These results conform to those of many previously studied SDVSAS (Konovalov and Ryzhkina, 2014). The 230nm average hydrodynamic diameter is typical for the supra- present in SDVSAS for C<Cthr (Yinnon and Liu, 2015c). DLS also evidenced associates in SDVSAS48, SDVSAS24 and SDVSAS12, but did not uncover any associates in SDVSAS6.4. The size distributions of the associates in SDVSAS48 and SDVSAS24 were mono-modal. These distributions were the same as those of SDVSASLB. However, decreasing the magnetic field’s amplitude from 24 to 12A/m led to poorly reproducible results, as the SDVSAS12 samples had poorly reproducible bimodal associate size distributions.

In the following, two complementary possible explanations for the findings of D.A. Konovalov et al. (2015) are offered:

♣ As noted in paragraphs I and V, interactions between UV EM fields and H2O underlie formation of , while interactions between IR EM fields and H2O underlie formation of CDrot. Despite Permalloy screening ambient UV and IR EM fields, the analyses of the experimental data obtained by Montagnier et al. (2009) have indicated that placing SDVSAS samples in Permalloy containers does not lead to immediate disintegration of all its molecular associates. In some samples, the associates seem to persist for at least 48 hours. Thus, D.A. Konovalov et al. (2015)’s finding that DLS measurements did not reveal associates in SDVSASP, likely does not mean that no associates were any longer present in samples after these were screened from ambient EM fields for 24 hours. It is conceivable that the screening considerably diminished the number of associates,  so that their prevalence was below the detection limit of DLS. According to QED, on disintegration of part of the supra-CDrot and supra-, the remaining domains are less stable. For , their destabilization implies that the continuing adsorption and desorption of their H2O (described in paragraph iii) take place on shorter timescales,  so that it becomes more difficult to observe these domains.

The difference between SDVSASP and SDVSASMF-7.85 is that the latter were exposed to 7.85 Hz EM fields. Since DLS showed that the size distribution of the domains in SDVSAS48 and SDVSAS24 was similar to that of SDVSASLB, i.e., these liquids contained nano-associates with a 230 nm average hydrodynamic diameter, the presence of the 7.85 EM field seemingly enhanced stabilization of . The stabilization is likely attributable to processes involving the vortices of the plasma of the QFE of These vortices are created by the 7.85 Hz EM field (see paragraph vii). For example, one possible stabilization process is that of the vortices of different  interacting with each other. Another process is that these vortices interact with the terrestrial magnetic field. This explanation implies that the intensity of the alternating 7.85 Hz EM field must be sufficiently strong to extract ions cycling around  from their orbit, because such extraction creates vortices (see paragraph vii). In other words, on lowering of the intensities of the applied 7.85 Hz field, fewer or no vortices are created in the plasma of QFE of . It is to be expected that few vortices cannot considerably stabilize , and that these cannot significantly reduce the rate of the continuous adsorption and desorption of the H2O of these domains. This explanation is compatible with the difficulties D.A. Konovalov et al. (2015) faced in observing molecular associates in SDVSAS12 . It is also compatible with them  not being able to discern associates in SDVSAS6.4.

In paragraph V it was emphasized thatn may get stabilized by CDrot. In other words, that the , which are meta-stable in bulk water at ambient conditions (see paragraph iii), get stabilized by CDrot. The stabilization of  in SDVSAS occurs when these liquids are diluted below and CDrot domains are stabilized by EDAIPDplasma or polar macromolecules. The findings of D.A. Konovalov et al. (2015) imply that SDVSAS stabilization of requires the presence of CDrot as well as the presence of 7.85 Hz magnetic fields. Indeed, the recent experiments by D’Emilia et al. (2015, 2016) hint that in bulk water, such fields reduce the continuously changing space distribution of the H2O constituting its meta-stable and its non-coherent H2O. Their findings suggest that exposing pure water to sufficiently strong magnetic fields with a frequency equaling that of one of the ion cyclotron frequencies of H3O+, e.g., vc=7.83 Hz, reduces the flickering of its landscape. In other words, it reduces the frequency of the adsorption and desorption of these domains’ H2O.

♣ Non-linear optical processes, induced by 7.85 Hz magnetic fields, are also possible sources for stabilizing the in SDVSAS48 and SDVSAS24. In other words, the 7.85 Hz EM field, the ULF radiation emitted from the QFE of , the UV radiation emitted from disintegrating , and the IR radiation emitted from disintegrating CDrot all may non-linearly interact with the molecules constituting these domains. These non-linear interactions may re-stabilize disintegrating and CDrot.  Since the H2O organized in the supra- oscillate coherently in phase with a coherently condensed EM field, and the same holds for the H2O constituting the supra-CDrot, non-linear optical processes may play significant roles. Coherence of photons is known to enhance non-linear optical effects (Zheltikov et al., 2007).

As to the likelihood of non-linear optical processes playing a role in stabilization of CDrot and , phenomena pertaining to the exclusion zone (EZ) water forming in water adjacent to a hydrophilic membrane should be mentioned. The analyses by Del Giudice et al. (2013) and Yinnon et al. (2016) have indicated that EZ water is composed of  and CDrot. The width of the EZ zone has been observed to expand on its irradiation with UV-visible and IR radiation, such as 270 – 700 nm and 1750 – 4200 nm radiation (Chai et al., 2009). On irradiation, up to 300 percent widening of the EZ has been observed. The degree of widening increased with wavelength, except for a few local maxima. These maxima appeared at wavelengths for which the absorption of radiation by the EZ is maximal, e.g., at 270 nm and 3.1 µm. The widening of the EZ was observed to be a function also of the irradiation time and intensity. These phenomena allude that non-linear optical processes affect the expansion of the EZ (i.e., affect the stabilization of and CDrot).

To investigate the role of non-linear optical processes in stabilization of and CDrot in SDVSAS, research is called for that focuses on excitation of these liquids not just by 7.85 Hz radiation but also with EM radiation at a large variety of frequencies.


Structural Aspects of SDVSAS Diluted Below the Avogadro Limit 

Recently, the group of A.I. Konovalov extended its DLS measurements to SDVSAS diluted below the Avogadro limit (Ryzhkina et al., 2015b). The group measured the DLS of SDVSAS of polyclonal antibodies to the protein interferon gamma. The SDVSAS were prepared with freshly double distilled water. Glass vials (Wiegand 120011543) were used. The stock solution was centesimally diluted. The concentration of the stock solution was 0.1 mg/ml. The SDVSAS samples were labeled as C2-C15, C30 and C50. The C indicates that the SDVSAS were centesimally diluted. The number denoted the number of dilutions steps (NDS). For the C7 samples, the molarity was calculated. It was of the order of 10-18 M. The properties of all the samples were measured after they were kept for 18 – 20 h on the laboratory bench or in Permalloy containers. The former are denoted as  (i.e., SDVSAS of anti-bodies kept at the laboratory bench). The latter are denoted as  ( i.e., SDVSAS of anti-bodies kept in Permalloy containers).

Ryzhkina et al.(2015b) also measured the electrical conductivity (χ), the surface tension (σ) and the ∆pH of   and . ∆pH is the difference between the pH of the SDVSAS and the pH of the double distilled water with which the SDVSAS were prepared, (the pH of the distilled water on the day the SDVSAS were prepared).

For controls, Ryzhkina et al.(2015b) prepared samples of serially diluted, vigorously shaken water (SDVSW), (double distilled water which was serially diluted with double distilled water and vigorously shaken after each dilution step).

A concise summary of the findings of Ryzhkina et al. (2015b) is presented in the next three paragraphs. Following the summary, these findings are analyzed within the context of the QED model of SDVSAS.

For  with concentrations above the Avogadro limit, just as Konovalov and Ryzhkina (2014) had observed for SDVSAS of many other solutes, Ryzhkina et al. (2015b) found that their χ, σ and ∆pH values vary non-monotonically with NDS. For  diluted below the Avogadro limit, they observed that these values also vary non-monotonically with NDS. Moreover, for SDVSW, they observed that the values of these variables vary non-monotonically with NDS and differ from those of  . In addition, they measured significant differences between the values of these variables for  or SDVSW versus those of double distilled water. [Elia and Niccoli (1999, 2000, 2004a) and Elia et al. (2005, 2015) did also report non-monotonic dependencies of χ, pH and other physicochemical properties of SDVSAS and SDVSW on NDS. They observed such non-monotonicity for SDVSAS of many different solutes. These SDVSAS had concentrations above the Avogadro limit or were diluted below this limit.]

For , Ryzhkina et al. (2015b) found that their χ and ∆pH variables only significantly vary non-monotonically with NDS when these liquids were less than 13 times centesimally diluted. When  and  were 13 times or more centesimally diluted, the χ and ∆pH of the former are higher or equal to those of the latter. As to σ, Ryzhkina et al. (2015b) observed an analogous behavior when  and  were more than 12 times centesimally diluted.

As to the DLS measurements of  and , the main results of Ryzhkina et al. (2015b) are:

A. For  with C2-C6, C8, C10-C12, C14 and C15, DLS does not indicate presence of associates.

B. For  with C7 or C9, DLS revealed presence of nano-associates with a monomodal size distribution. The average diameters of the associates are respectively, ~170 nm and ~140 nm.

C. The associates in C7  are absent in C7 .

D. For C9  and C9 their associates have nearly the same dimensions. The monomodal size distribution of is a bit narrower than that of  .

E. In C13, C30 or C50 , as well as in C13, C30 or C50  , DLS revealed particles with a trimodal size distribution. For such a distribution, the DLS method is inadequate for accurately measuring the average sizes of the particles. Yet the DLS measurements hint that particles with sizes of the order of 10-9 m, 10-8 m and 10-7 – 10-6 m are present in these liquids.

For analyzing the above cited findings of Ryzhkina et al. (2015b), foremost, the following should be emphasized: although DLS does not indicate presence of associates in C2 -C7, C8, C10-C12, C14 and C15 of  (see paragraph A), these liquids’ physicochemical variables and the non-monotonic variance of these variables with NDS reflect that these liquids’ properties differ from those of double distilled water. Analyses of experimental data reported in earlier publications, pertaining to measured physicochemical variables of SDVSAS of other solutes with concentrations above the Avogadro limit or which were diluted below this limit (Elia and Niccoli, 1999, 2000, 2004a; Elia et al., 2005, 2015, 2008a&b; Belon et al.2008), as well as those pertaining to SDVSW, suggest that for  C < = Cthr, in these liquids, CDrot underlie the non-monotonic variance of their physicochemical variables (Yinnon and Yinnon, 2011; Yinnon and Elia, 2013; Yinnon and Liu, 2015b&c). The analyses indicate that the impact of CDrot on the physicochemical properties of SDVSAS or SDVSW slowly increases with the age (months, years) of these liquids. Since the data show that the physicochemical variables greatly vary with the age of those SDVSAS which were only a few times diluted, shortly shaken and stored for less than a year, it suggests the following: the sizes of the CDrot in such “young” SDVSAS are still small and the ordering of these domains in supra-CDrot vary greatly with time. Based on these analyses, the findings presented in paragraph A hint that DLS is insensitive to the presence of young CDrot. This insensitivity will be discussed further on.

As to the findings presented in paragraphs B and C, monomodal distributions of molecular associates with average diameters of the order of 10-7 m, which disappeared on screening of the liquid by Permalloy, (like the one observed for C7 ) have been detected in SDVSAS of many solutes (Konovalov and Ryzhkina, 2014). The associates underlying such distributions were shown to have the typical characteristics of  and supra- (Yinnon and Liu, 2015b&c). As noted in previous sections, for C < Cthr, these domains are stabilized by CDrot and low frequency magnetic fields. It is thus most likely that the 170 nm sized associates observed in C7  are supra-. Measurements of their electrokinetic potential are called for. Such measurements could provide additional evidence for this conclusion.

A question arising from the above presented data is: “Why did not DLS detect a monomodal distribution of associates with sizes of the order of 10-7 m in  which were less than 7 times centesimally diluted (see paragraph A)?” The concentrations of the antibodies in C2 – C7  are in the range of 10-8 – 10-18 M. A value as low as about 10-18 M has not been reported for  = Cthr of SDVSAS of other solutes. Typically, ~10-10 M <  = Cthr< ~10-6 M.

A possible answer is that in  which were less than 7 times centesimally diluted, the alignment of CDrot in supra-CDrot and the incorporation of the solutes in these domains are detrimental to stabilization of  in numbers sufficient for these domains to be observable with DLS. Such an answer might also be applicable to C8, C10-C12, C14 and C15 , in which as noted in paragraph A, DLS did not detect associates with diameters of the order of 10-7 m. A similar explanation has been given for the relation between the prevalence of molecular associates with sizes of the order of 10-7 m observed in SDVSAS of many solutes and the dielectric permittivity of these liquids (Konovalov and Ryzhkina, 2014; Yinnon and Liu, 2015c). (The dielectric permittivity reflects the alignment of the electric dipoles of CDrot in supra-CDrot.)

The above analyses raise the issue: “How is information on CDrot and supra-CDrot reflected in DLS data?” At least theoretically, these domains may grow to sizes of 10-5 – 10-4 m. The issue has not been clarified in the earlier published analyses of SDVSAS. Some clues are provided by atomic force measurements (AFM) of the residues left over after evaporating drops of SDVSAS. AFM data revealed that in addition to more or less spherical 10-7 m sized chunks, micron sized elongated molecular associates were present in these residues [see Fig. 3f in Ryzhkina et al. (2012b); Fig. 1 in Ho (2014); Fig. 5 in Elia et al. (2015)]. The widths of these elongated associates were of the order of 10-9 – 10-7 m. Their lengths were of the order of 10-6 – 10-5 m. These elongated associates were more or less linear or curled up. The diameters of the ball-shaped curled up ones were of the order of 10-7 – 10-6 m. Earlier published analyses of the various physicochemical properties of these SDVSAS evoked that the elongated associates are CDrot, or supra-CDrot organized in chains with their electric dipole moments parallel oriented (Yinnon and Liu, 2015c). The analyses also indicated that the chunks are supra-. Therefore, in relation to the abovementioned issue, it is my conjecture that DLS cannot clearly distinguish between the presence of supra- and the presence of the supra-CDrot when these elongated associates are curled up in balls. Moreover, it is likely that DLS is insensitive to very thin linear molecular associates, such as the more or less linearly micron sized CDrot or supra-CDrot with width of the order of 10-9 m.

In regard of this conjecture, the data reported in paragraphs B and D possibly can be understood as follows: The monomodal distribution of associates with sizes of the order of 10-7 m which were present in C9  mainly reflect the presence of curled up supra-CDrot and few . On screening of this  by Permalloy for 18 – 20 h, most of these few  disintegrated. However, as noted above, CDrot have much slower dynamics than Therefore, CDrot disintegrate slower than . So many supra-CDrot could persist in the samples after their storage for 18 – 20 hours in Permalloy containers. Moreover, since CDrot formation is mediated by far IR radiation, while  formation is mediated by UV radiation, it is possible that screening by Permalloy affects the former less than the latter. These explanations potentially clarify why the difference between the observed DLS results for C9 and C9 is small.

As to the findings presented in paragraph E, my conjecture also enables explanation of the characteristics of the associates in C13, C30 or C50 of  and . The associates with diameters of the order of 10-7 – 10-6 m, which were revealed in these  by DLS, likely mainly are curled up CDrot and supra-CDrot. These CDrot may have stabilized few 10-7 m sized . On screening these  with Permalloy for 18 – 20 hours, many of the CDrot prevailed, while the mostly disintegrated. The persistence of CDrot explains that the χ, σ and ∆pH values of these significantly differed from those of double distilled water.

As to the 10-9 m and 10-8 m sized particles which were observed in C13, C30 or C50 of  and in , these may include broken CDrot pieces (i.e., EDACDrot) created by the vigorous shaking of these liquids. The slow dynamics of CDrot implies that their size distribution is wide. The 10-9 m sized particles also may be (hydrated) compounds released by the containers or other impurities. Such particles were not discerned with DLS in or in diluted less than 13 times. Likely, in these solutions, their concentrations were very low, i.e., too low to be detectable with DLS. The concentration of compounds released by the glass container slowly increases with the number of dilution steps, because after each dilution step the liquid is vigorously shaken and a percentage of the liquid is transferred to a new vial containing double distilled water.

The upshot of the above analyses is that in the , which were diluted below the Avogadro limit, CDrot and supra-CDrot are apparently present. Also analyses reported in earlier publications pertaining to SDVSAS of other solutes which were diluted below this limit, as well as those pertaining to SDVSW, suggested presence of such domains in these liquids (Yinnon and Elia, 2013). However, the analyses presented in the previous paragraphs hint that the capability of DLS to detect these domains depends on their shape. Since experimental data hint that on aging of SDVSAS, CDrot grow and organize in supra-CDrot (Yinnon and Elia, 2013), DLS measurements of aged SDVSAS are called for. In addition to DLS, it seems desirable to employ other techniques for revealing typifying fingerprints of CDrot and supra-CDrot in SDVSAS diluted below the Avogadro limit (e.g., their ferroelectric orderings and the electric dipole rotations of their H2O.) Their ferroelectric orderings may be revealed by dielcometric titrations, as the experimental data obtained by Ryzhkina et al. (2012b, 2013) and their analyses by Yinnon and Liu (2015c) have indicated. Moreover, as the recent results obtained by Mahata (2013) suggest, such orderings and rotations of the electric dipoles of their molecules may possibly be revealed by dielectric dispersion measurements.

Dielectric dispersion measurements have provided evidence for the presence of about 3×10-5 m sized molecular structures in water samples to which a few drops of one of the following liquids were added: 6 or 30 times serially diluted solutions of Graphites or Copper, vigorously shaken after each dilution step (Mahata, 2013). The solvent of these serially diluted solutions was ethanol. After addition of the drops to the water samples, the samples contained about 3 percent ethanol. The 3×10-5 m sized structures affected the polarization of the samples. The effect became distinguishable about 24 – 48 hours after the drops were added to the water samples. The detailed experimental results indicate that formation of the structures involves electric dipole rotations of the liquids’ molecules. More accurately, it involves rotations of the solvent molecules, but does not involve rotations of Copper or Graphite molecules. The QED model of serially diluted, vigorously shaken polar liquids indicates that such liquids should contain CDrot (Yinnon and Yinnon, 2011). Analyses of experimental data pertaining to such liquids revealed fingerprints of such domains (Yinnon and Liu, 2015c). So the drops which Mahata (2013) added to the water samples probably contained CDrot composed of the polar ethanol molecules. Since such CDrot are aggregates with an electric dipole moment, QED predicts that these may stabilize CDrot composed of H2O in the water samples (Del Giudice et al., 1988). So it is likely that the 3×10-5 m sized molecular structures are CDrot. This is likely because according to QED, electric dipole rotations underlie the CDrot composed of ethanol or water. However, it should be emphasized that Mahata (2013) observed the structures in water containing about 3 percent ethanol. In such aqueous solutions, as when neither the water nor the ethanol are serially diluted and vigorously shaken, domains are present (Sedlak, 2006; Sedlak and Rak, 2013). These domains have characteristics of QED coherent domains (Yinnon and Yinnon, 2009). Recent analyses indicated that in water containing a few percentage of ethanol, fractal structured molecular associates are present (Ghosh and Bagchi, 2016). As noted above, fractality is a property of supra-QED coherent domainsf. So repeating the experiments of Mahata (2013) for controls is called for. For example, the controls may be water samples to which drops of SDVSAS of Copper or Graphite were added — according to the above and previous published analyses, these drops should contain CDrot.  Also, water samples to which drops of non-serially diluted non-vigorously shaken ethanol solutions of Copper or Graphite are added may serve as controls — these liquids should not contain CDrot. Yet, even in the absence of data for controls, the findings of Mahata (2013) are significant. They suggest that dielectric dispersion measurements for SDVSAS diluted below the Avogadro limit seem to be capable of providing more direct evidence of CDrot in these liquids.

 

Discussion

The above analyses show that the QED model of SDVSAS explains the recently observed self-similarity topology of these liquids. This self-similarity, together with the dissipative properties of SDVSAS identified in previous studies (Elia et al., 2000, 2008a&b; Belon et al., 2008; Elia et al., 2015), in combination with the analyses by Vitiello (2012, 2014), corroborates that QED coherence plays a significant role in these liquids.f In other words, the above analyses substantiate the model’s assumption that QED coherence domains (e.g., CDrot, and IPDplasma), underlie the structural properties of SDVSAS.

Furthermore, the above analyses show that the QED model of SDVSAS clarifies the EM phenomena of these liquids. The model elucidates some missing links in the explanations of Montagnier et al. (2011, 2015) pertaining to the emission of ULF radiation by SDVSAS of DNA.

The above analyses also complement previous ones, which showed that the QED model of SDVSAS consistently explains many structural and physicochemical properties of these liquids (Yinnon and Yinnon, 2011; Yinnon and Elia, 2013; Yinnon and Liu, 2015a-c). The above analyses of the recent DLS, electrical conductivity, pH and surface tension measurements carried out by Ryzhkina et al. (2015b) support the model. These recent measurements, which pertain to SDVSAS diluted beyond the Avogadro limit, complement the many earlier measurements for SDVSAS diluted up to this limit. The analyses indicate that the nano-associates observed in SDVSAS diluted below the Avogadro limit are also QED coherence domains, i.e., CDrot and .
The analyses presented above, together with those presented in earlier publications and schematically portrayed in the lecture by Konovalov et al. (2014)g and the poster by Yinnon and Konovalov (2014)h, accentuate that an appropriate description for the molecular associates in SDVSAS is the term recently coined by Konovalov and Ryzhkina (2016): “nano-sized self-organized substrate-induced molecular ensembles.” The term “substrate” here refers to the solute with which the SDVSAS is prepared. At concentrations slightly lower than the transition concentrations or , the solute molecules serve as substrates which may stabilize respectively, IPDplasma or CDrot. Vigorous shaking of the liquid excites the domains, i.e., creates electric dipole aggregates.  For C < , these aggregates stabilize additional CDrot. Thus for C < , the creation of new electric dipole aggregates no longer depends on the solute molecules. The CDrot may stabilize . At C ~ or at lower C, the solute properties become imprinted in the CDrot and . On diluting the liquid, part of these domains get transferred to the dilute. Vigorous shaking of the dilute breaks up the CDrot and thus creates new electric dipole aggregates. These aggregates stabilize new CDrot and . At C above the Avogadro limit but below , the solutes locate within CDrot and may orbit around . At such low concentrations, solutes still may affect physical, chemical or biological processes, e.g., through resonances (Del Giudice et al., 2010b). For SDVSAS diluted below the Avogadro limit, the current data hint that the physicochemical and biological properties of these liquids are attributable to CDrot, and the imprint of the characteristics of the substrates (original solutes) in these domains.

f See the discussions of Vitiello (2012, 2014), respectively, on “Fractals, Dissipation and Coherent States” and on “The isomorphism between dissipative systems, fractal self-similarity and electro-dynamics.”

g Konovalov A. I., Ryzhkina  I. S., Yinnon T. A. (2014).  Molecular assemblies mediated by electromagnetic fields in highly diluted aqueous solutions: Accordance of experimental data and predicted by electro-dynamic theory data. VIIth International Symposium  “Design and Synthesis of  Supramolecular Architectures, Kazan, Russia; Annual Conference on the Physics, Chemistry and Biology of Water, Bulgaria — http://www.waterconf.org/participants-materials/2014/Presentations/Alexander_Konovalov_Water_Conference_2014.pdf

h Yinnon T. A. and Konovalov A. I. (2014). Domains formation mediated by electromagnetic fields in dilute solutions. Annual Conference on the Physics, Chemistry and Biology of Water, Bulgaria — see Tamar Yinnon Research Gate.

The above analyses support the following conclusions, which recently were drawn by Konovalov and Ryzhkina (2016): At concentrations below Cthr but above the Avogadro limit, the characteristics of the nano-associates imply that these associates constitute a phase which is different from the medium, i.e., SDVSAS are nano-heterogeneous. Consequently, these SDVSAS cannot strictly be regarded as solutions in the customary sense. Instead, these are nano-disperse systems. Nano-associates composed of H2O constitute the disperse phase. According to QED, although not yet experimentally verified, the solute (substrate) molecules locate in the disperse phase. On serial diluting these systems and vigorous shaking after each dilution step, the solute molecules decrease in numbers and ultimately these are not any longer present in the liquid. However, even beyond the Avogadro limit, nano-associates composed of H2O persist. Thus on serial diluting a SDVSAS beyond the Avogadro limit, in combination with vigorous shaking the liquid after each dilution step, a special kind of disperse system emerges — a disperse system of “water in water.” The disperse phase consists of nano-associates composed of H2O. The electronic structure and the ordering of the H2O in the associates are determined by the characteristics of the substrate molecules. Therefore, the associates contain the molecular information of the substrate.

The above analyses in combination with previous ones show that the QED model of SDVSAS adequately describes many aspects of these liquids. A shortfall of the model is that it does not specify the relation between the characteristics of the substrates and the physicochemical, structural and bioactive properties of these liquids. Future research directed at such derivations is called for. The challenges involved are not just theoretical. The many experimental studies of SDVSAS hitherto have shed very little light on this relation. For highly diluted SDVSAS, until now, UV, IR, ULF or NMR spectra did not reveal any solute specific effects on their molecules’ oscillations or on the structure of their associates. The UV absorbance and fluorescence spectra of SDVSAS prepared from different substrates may vary. However, no specific features attributable to a particular substrate have been identified (Lo,1996; Lobyshev et al., 2005; Wolf et al., 2011; Klein et al., 2013). The spectra consist of broad rather featureless bands. Also in IR and NMR spectra of SDVSAS, no solute specific features have yet been observed (Zubareva et al., 2003a&b; Lo et al., 2009; Demangeat, 2009, 2010; Elia et al., 2015). The group of Elia et al. (2012) demonstrated that aqueous SDVSAS emit radiation. They demonstrated this for SDVSAS of fullerene which were 5 to 30 times centesimally diluted and for SDVSAS of carbon nanotubes which were 5 times centesimally diluted. Both these SDVSAS were prepared in glass vessels. However, their analyses did not reveal the impact of the original solute on the emitted radiation. As to ULF emission by SDVSAS, the currently available data are very limited. The data hint that only for certain types of DNA the emission is sufficiently strong to be observable with the method developed by Montagnier et al. (2009). The effects of excitation of SDVSAS by 7.85 Hz EM field, observed by D.A. Konovalov et al. (2015) and analyzed above, suggest that various non-linear optical processes occur in these liquids. Such processes pose challenges for revealing details of highly diluted SDVSAS with absorption or fluorescence spectroscopy.

So far, to the best of my knowledge, the sole significant findings pertaining to the relation between substrates’ characteristics and the physicochemical and structural properties of SDVSAS are:

✥ The bioactivity, physicochemical and structural properties of SDVSAS are interrelated (Lobyshev et al., 2005; Palmina et al., 2009; Ryzhkina et al., 2010, 2011b, 2012b, 2013; Konovalov and Ryzhkina, 2014). The bioactivity is solute specific.

✥  The thermoluminescence spectra of 10-30 g/cm3 SDVSAS of LiCl and SDVSAS of NaCl differ (Rey, 2003). The difference pertains to the hydrogen bond network in these SDVSAS. The hydrogen bond network in SDVSAS of LiCl is more suppressed than that in SDVSAS of NaCl. Electro-chemical research has shown that Li+ ions attract more hydration H2O than Na+ ions, because the former have a smaller radius (Robinson and Stokes, 2002). As a result, the hydrogen-bond suppressing effect of LiCl salts is larger than that of NaCl salts. Thus, the thermoluminescence measurements revealed that the hydrogen bond suppressing effect of these salts even persists in their SDVSAS diluted to the Avogadro limit.

Uncovering the relation between the substrates’ characteristics and physicochemical and structural properties of highly diluted SDVSAS is important for identifying the processes underlying their bioactivity. Several hypotheses concerning the active principle in SDVSAS have been forwarded:

♦ Yinnon and Liu (2015c) surmised that the  present in SDVSAS constitute the active principle, e.g., that the electronic structure of the substrate leaves an imprint on the pool of quasi-free electrons (QFE) of . The substrate affects the vortices in the QFE which are induced by the vigorous shaking of the SDVSAS. Moreover, they conjectured that the polarity of the substrate affects the ordering of CDrot. They based their conjecture on the experimental findings of Konovalov and Ryzhkina (2014) – for SDVSAS with ~10-18 M<C<Cthr, their bioactivity is related to the hydrodynamic diameter and the electrokinetic potential of their molecular associates with sizes of the order of 10-7 m. Yinnon and Liu (2015c) showed that these associates have the typical properties of .  Based on their analyses, they surmised that these domains affect electron transport processes. Experiments have indicated that SDVSAS of Melafen [the melamine salt of bis (hydroxymethyl) phosphinic acid dihydrate] affect electron transport in the mitochondrial respiratory chain, even when the concentration of these solutions is as low as 10-18 – 10-22 M (Zhigacheva et al., 2009).

♦ Chakraborty et al. (2015) and Bell et al. (2015a) have surmised that the bioactivity of SDVSAS is related to traces of the original solute molecules. These molecules may organize in structures composed of compounds released by the containers. Traces of solutes in 30 and 200 times centesimally serially diluted, vigorously shaken solutions of several metals, with ethanol as solvent, have been reported (Chikramane et al., 2012). The solutions were prepared by pharmaceutical companies. However, a follow up study by Bell et al. (2015b) showed that inductively-coupled plasma mass spectroscopy could not detect traces of silver in 30 and 200 times centesimally serially diluted, vigorously shaken solutions of silver with ethanol as solvent. These studies and others also detected presence of compounds released by the glass containers, their cork stoppers and other impurities in extremely diluted solutions (Bell et al., 2015c; Witt et al., 2006; Elia and Napoli, 2010).

It is, however, doubtful that traces of the original solute could be the solute specific bioactive element in SDVSAS, as hinted by the following experimental results. As shown by Konovalov and Ryzhkina (2014), for SDVSAS with 10-18 M <C<Cthr, their solute specific bioactivity disappeared when samples were stored under hypo-electromagnetic conditions,  (in Permalloy containers). Under such storage, their 10-7 m and 10-6-10-5 m sized associates, which mainly are composed of H2O, also disappeared. Moreover, after such storage, the physicochemical properties of SDVSAS returned to those of serially diluted solutions which were not vigorously shaken after each dilution step. In other words, after storage in Permalloy containers, neither the very tiny amounts of solutes still present in solutions with 10-18 M <C<Cthr, nor the compounds released by the containers or other impurities, could trigger specific bioprocesses. Furthermore, Montagnier et al. (2011) demonstrated the following for a sample of SDVSAS of DNA kept in a closed plastic Eppendorf tube and a sample of SDVSW kept in a closed plastic Eppendorf tube, which were both placed in the same Permalloy container and exposed for 18 h to a very low frequency (7 Hz) magnetic field: The ULF radiation emitted by the SDVSAS induced solute specific properties in the SDVSW. In other words, even when there was no direct contact between the SDVSAS samples and the SDVSW samples, they had similar properties. Additionally, Davenas et al. (1988) showed that for serially diluted, vigorously shaken solutions of anti-IgE antibody  diluted below the Avogadro limit, their bioactivity varied quasi-periodically with the number of dilution steps (NDS). It is hard to envision, within the context of the customary electro-static theories, that compounds released by the containers or other impurities can induce such quasi-periodic patterns in liquids.

Very recent findings raise another hypothesis for the bioactivity of highly diluted SDVSAS. Capolupo et al. (2016) showed that for DNA and enzymes embedded in water, the electric dipole-dipole interactions between the aromatic ring structures in DNA and the enzymes may get correlated by the QED interactions between rotating H2O. In other words, the DNA radiative dipole field triggers collective dipole waves in its surrounding water, i.e., stabilizes CDrot, which in turn couple with the enzyme’s radiative dipole field. Based on these findings by Capolupo et al. (2016), it is plausible that by adding SDVSAS to the DNA-enzyme-water system, the CDrot present in the former affect the CDrot present in the latter. For SDVSAS, the relation between the substrate and the ferroelectric ordering of the H2O constituting their CDrot have not yet been derived. However, analogous to the imprint of information in magnetic materials, the substrates’ properties can affect the ferroelectric orderings of CDrot. Some experimental findings provide support for this hypothesis. Belov et al. (2004) showed that when only polar solvents were used for preparing serially diluted, vigorously shaken liquids with concentrations below 10-18 M and and even when such liquids were diluted beyond the Avogadro limit, these could affect microsomal membranes. CDrot only form in polar liquids. It is thus possible that the ferroelectric ordering of the solvent molecules constituting CDrot affects membranes. However, it is also possible that the presence of CDrot in serially diluted, vigorously shaken polar liquids is required just for stabilizing the domains consisting of solvent molecules, which coherently transit between two of their electronic states, such as the in water and their analogues in other polar liquids. In case the sole function of CDrot is stabilization of , the latter can be regarded as the unit carrying the substrates’ information relevant for bio-processes.

 

Summary and Conclusions

The importance of SDVSAS for toxicology, pharmacology and agriculture mandates research directed at uncovering their bioactive principle. During the last decades, much information has been gathered on the physicochemical and structural properties of SDVSAS. For such properties that still require additional clarification, some warranted research projects have been specified in the previous sections. During the last decades, much information has also been gathered on the physicochemical and structural properties of serially diluted solutions, which were prepared with ethanol or other polar solvents instead of water. The QED models of all these solutions are very similar. The models provide consistent explanations for the liquids’ physicochemical and structural properties, as shown in previous publications and supplemented in the current paper.

Experimental data and QED theory, however, leave us in the dark in regards to the relation between the bioactivity of the aforementioned solutions and their physicochemical properties. Experimental data and QED theory also leave us in the dark about the relation between the characteristics of the substrates (solutes) and the physicochemical properties of the aforementioned solutions. Therefore, research projects are called for that aim to develop techniques capable for elucidating the relation between the substrates’ characteristics and the structure, physicochemical and bioactive properties of highly diluted SDVSAS. At the current stage of knowledge, only few vague road signs are available for directing such research (e.g., electron transport processes may be involved.) Yet, the success of the QED model for elucidating the physicochemical and structural properties of serially diluted polar liquids indicates that it can serve as a tool for identifying techniques which can reveal the information of the substrate’s imprint in these liquids.

 

Acknowledgements

With appreciation, I thank Prof. A. I. Konovalov for reading the manuscript and his constructive comments. I express my gratitude to Prof. A. M. Yinnon for his continuous support, encouragement and proofreading of the manuscript. I also express my appreciation and thankfulness to Jesús Aguilar for his extensive literature search.

References

Ambrosetti A, Ferri N, DiStasio RA Jr., Tkatchenko A (2016). Wavelike charge density fluctuations and van der Waals interactions at the nanoscale. Science 351:1171-1176.

Arani R, Bono I, Del Giudice E; Preparata G (1995). QED coherence and the thermodynamics of water. Int J Mod Phys B 9: 1813-1841.

Bell IR, Schwartz GE, Frye J, Sarter B, Standish LJ (2015a) Extending the Adaptive Network Nanomedicine Model for Homeopathic Medicines: Nanostructures as Salient Cell Danger Signals for Adaptation. Nanosci Technol 2: 1-22.

Bell IR, Muralidharan S, Schwartz GE (2015b). Nanoparticle Characterization of Traditional Homeopathically-Manufactured Silver (Argentum Metallicum) Medicines and Placebo Controls. J Nanomed Nanotechnol 6:311.

Bell IR, Muralidharan S, Schwartz GE (2015c). Nanoparticle Characterization of Traditional Homeopathically-Manufactured Gelsemium sempervirens Medicines and Placebo Controls. J Nanomedine Biotherapeutic Discov 5:136.

Belon P, Elia V, Elia L, Montanino M, Napoli E, Niccoli M (2008). Conductometric and calorimetric studies of the diluted and agitated solutions. On the combined anomalous effect of time and volume parameters. J Therm Anal Calorimetry 93:459-469.

Belov VV, Mal’tseva EL, Pal’mina NP, EB Burlakova (2004). The Role of the Solvent Polarity in the Mechanism of Action of Biologically Active Compounds at Ultralow Concentrations. Dokl Biochem Biphys 399:362-364.

Belov VV, Mal’tseva EL, Pal’mina NP (2011). Modification of the structure of plasmatic membranes of the liver by the action of α-tocopherol in vitro. Biophysics 56:323-330.

Bono I, Del Giudice E, Gamberale L, Henry M (2012). Emergence of the Coherent Structure of Liquid Water.  Water 4: 510-532.

Burlakova EB, Grechenko TN, Sokolov EN, Terekhova SF (1986). Effect of inhibitors of radical reactions of lipid oxidation on electrical activity of the isolated snail neuron. Biofizika 31:921-923.

Burlakova EB, Konradov AA, Maltseva EL (2004). Effect of extremely weak chemical and physical stimuli on biological systems. Biophysica (Moscow) 49:522-534.

Burlakova EB (2005). Bioantioxidants: yesterday, today, tomorrow. Pages 18-27 in Chemical and biological kinetics – New Horizons. Vol. 2. Biological kinetics. Eds. Burlakova EB, Shilov AE, Varfolomeev SD, Zaikov G. Koniklijke Brill NV, Leiden, The Netherlands.

Capolupo A, Del Giudice E, Elia V, Germano R, Napoli E, Niccoli M, Tedeschi A, Vitiello G (2014).  Self-similarity properties of Nafionized and filtered water and deformed coherent states. Int J Mo. Phys. B 28, 1450007.

Capolupo A, Craddock TJA, Kurian P, Vitiello G (2016). Water-mediated correlations in DNA-enzyme interactions. arXiv:1608.08097[physics.bio-ph].

Cabral do Couto P, Chipman DM (2012). Insights into the ultraviolet spectrum of liquid water from model calculations: The different roles of donor and acceptor hydrogen bonds in water pentamers. J Chem Phys 137:18430.

Celghini E, Graziano E, Vitiello G (1990). Classical Limit and Spontaneous Breakdown of Symmetry as an Environment Effect of Quantum Field Theory. Phys Lett A 145:1-6.

Chai B-H, Zheng J-M, Zhao Q, Pollack GH (2008). Spectroscopic studies of solutes in aqueous solution. J Phys Chem A 112:2242-2247.

Chai B-H, Yoo H, Zhao Q, Pollack GH (2009). Effect of Radiant Energy on Near-Surface Water. J Phys Chem B 113: 13953–13958.

Chakraborty M, Ghosh S, Das S, Basu R, Nandy P (2015). Effect of Different Potencies of Nanomedi-cine Aconitum Napelles on Its Spectral and Antibacterial Properties. Int J Innov Res Sci Eng Technol 4:6861-6867.

Chikramane PS, Kalita D, Suresh AK, Kane SG, Bellare JR (2012). Why Extreme Dilutions Reach Non-zero Asymptotes: A Nanoparticulate Hypothesis Based on Froth Flotation. Langmuir 28: 15864-15875.

Ciavatta L, Elia V, Napoli E, Niccoli M (2008). New physico-chemical properties of extremely diluted solutions. Electromotive force measurements of Galvanic cells sensible to the activity of NaCl at 25 oC. J Solution Chem 37:1037-1049.

Cifra M (2015). Cellular electrodynamics in kHz–THz region. D. Fels, M. Cifra and F. Scholkmann (Eds), Fields of the Cell, p. 189–214.

Davenas E, Beavais F, Amara J, Oberbaum M, Robinzon B, Miadonna A, Tedeschi A, Pomeranz B, Fortner P, Belon P, Sainte-Laudy J, Poitevin B, Benveniste J (1988). Human basophil degranu-lation triggered by very dilute antiserum against IgE. Nature 333: 816-818.

Del Giudice E, Preparata G, Vitiello G (1988). Water as a free electric dipole laser. Phys Rev Lett 61: 1085-1088.

Del Giudice E, Preparata G (1998). “A new QED picture of water,” in Macroscopic Quantum Coherence, eds. Sassaroli E, Srivastava YN, Swain J, Widom A (World Scientific, Singapore).

Del Giudice E, Preparata G, Fleischmann M (2000). QED coherence and electrolyte solutions. J Electroanal Chem 482: 110–116.

Del Giudice E, Fleischmann M, Preparata G and Talpo G (2002). On the “unreasonable” effects of ELF magnetic fields upon a system of ions. Bioelectromagn. 23: 522–530.

Del Giudice E, Vitiello G. (2006). Role of the electromagnetic field in the formation of domains in the process of symmetry-breaking phase transitions. Phys Rev A 74: 022105-1-022105-9.

Del Giudice E (2007). Quantum Field Theory applied to the water structure. Summer School 2007 of the International Institute of Biophysics Neuss Germany.

Del Giudice E, Fuchs EC Vitiello G (2010a). Collective Molecular Dynamics of a Floating Water Bridge. Water 2: 69-82.

Del Giudice E, Spinetti PR and Tedeschi A (2010b). Water dynamics at the root of metamorphosis in living organisms. Water 2: 566-586.

Del Giudice E, Tedeschi A, Vitiello G, Voeikov V (2013). Coherent structures in liquid water close to hydrophilic surfaces. J Phys:Conf Ser442:012028.

Demangeat JL (2009). NMR water proton relaxation in unheated and heated ultrahigh aqueous dilutions of histamine: Evidence for an air-dependent supramolecular organization of water. J Mol Liq 144:32-39.

Demangeat JL (2010). NMR relaxation evidence for solute-induced nanosized superstructures in ultramolecular aqueous dilutions of silica–lactose. J Mol Liq 155:71-79.

D’Emilia E, Giuliani L, Lisi A, Ledda M, Grimaldi S, Montagnier L, Liboff AR (2015). Lorentz force in water: evidence that hydronium cyclotron resonance enhances polymorphism. Electromagn Biol Med 34: 370-375.

D’Emilia E, Ledda M, Foletti A, Lisi A, Giuliani L, Grimaldi S, Liboff AR (2017). Weak-field H3O+ ion cyclotron resonance alters water refractive index. Electromagn Biol Med 36:55-62.

De Ninno A, Congiu Castellano A (2011). Co-ordination of water molecules in isotopic mixtures. J Mol Struct 1006: 434-440.

Duval E, Adichtchev S, Sirotkin S, Mermet A (2012). Long-lived submicrometric bubbles in very diluted alkali halide water solutions. PhysChemChemPhys; 14: 4125-4132.

Elia V, Niccoli M (1999). Thermodynamics of Extremely Diluted Aqueous solutions. Ann New York Acad Sci 870: 241-248.

Elia V, Niccoli M (2000). New Physico-chemical properties of water induced by mechanical treatments: A calorimetric study at 25oC.  J Therm Anal Cal 61: 527-537.

Elia V, Niccoli M (2004a). New Physico-chemical properties of extremely diluted aqueous solutions. J Therm Anal Cal 75: 815-836.

Elia V, Napoli E, Niccoli M, Nonatelli L, Ramaglia A, Ventimiglia E (2004b). New Physico-chemical properties of extremely diluted aqueous solutions. A calorimetric and conductivity study at 25°C. J Therm Anal Cal 78: 331–342.

Elia, V, Marchese M, Montanino M, Napoli E, Niccoli M, Nonatelli L, Ramaglia A (2005). Hydrohysteretic phenomena of extremely diluted solutions induced by mechanical treatments: A calorimetric and conductometric study at 25 °C. J. Solut. Chem. 34:947–960.

Elia V, Napoli E, Niccoli M, Marchettini N, Tiezzi E (2008a). New physico-chemical properties of extremely dilute solutions. A conductivity study at 25°C in relation to ageing. J Solution Chem 37: 85-96.

Elia V, Elia L, Marchettini N, Napoli E, Niccoli M,Tiezzi E (2008b). Physico-chemical properties of aqueous extremely diluted solutions in relation to ageing. J Therm Anal Calorimetry 93:1003-1011.

Elia V, Napoli E. (2010). Dissipative structures in extremely diluted solutions of homeopathic medicines. A molecular model based on physico-chemical and gravimetric evidences. Int J Des Nat 5: 39-48.

Elia, V, Marrari LA, Napoli E. (2012). Aqueous nanostructures in water induced by electromagnetic fields emitted by EDS. J Therm Anal Calorim 107:843–851.

Elia V, Ausanio G, De Ninno A, Gentile F, Germano R, Napoli E, Niccoli M (2013a). Experimental evidence of stable aggregates of water at room temperature and normal pressure after iterative contact with a Nafion® polymer membrane. Water 5:16-26.

Elia V, Elia L, Napoli E, Niccoli M (2013b). Physical-chemical study of water in contact with a hydrophilic polymer: Nafion. J Therm Anal Calorim 112, 937-944.

Elia V, Lista L, Napoli E, Niccoli M (2014a). A Thermodynamic Characterization of Aqueous Nanostructures of Water Molecules Formed by Prolonged Contact with the Hydrophilic Polymer Nafion. J. Therm. Anal. Calorim. 115:1841-1849.

Elia V, Ausanio G, De Ninno A, Germano R, Napoli E, Niccoli M (2014b). Experimental Evidences of Stable Water Nanostructures at Standard Pressure and Temperature Obtained by Iterative Filtration. Water 5:121-130.

Elia V, Ausanio G, Germano R, Gentile F, Germano R, Napoli E, Niccoli M (2014c). Experimental Evidences of Stable Water Nanostructures in Extremely Dilute Solutions, at Standard Pressure and Temperatures. Homeopathy 103:44-50.

Elia V, Germano R, Napoli E (2015). Permanent Dissipative Structures in Water: The Matrix of Life? Experimental Evidences and their Quantum Origin. Curr Top Med Chem 15: 559-571.

Elia V, Yinnon TA, Oliva R, Napoli E, Germano R, Bobba F, Amoresano A (2017). Chiral micron-sized H2O aggregates in water: Circular dichroism of supramolecular H2O architectures created by perturbing pure water. Water 8:1-29.

Ferri N, DiStasio RA, Ambrosetti A, Car R, Tkatchenko A (2015). Electronic properties of molecules and surfaces with a self-consistent interatomic van der Waals density functional. Phys Rev Lett 114: 176802.

Fiorini RA (2016). Arbitrary multi-scale (AMS) systems biology and biomedical engineering effective modeling and simulation. International Journal of Biology and Biomedical Engineering (BIBIEN) 10:61-71.

Georgalis G, Kierzek AM, Saenger W (2000). Cluster Formation in Aqueous Electrolyte Solutions Observed by Dynamic Light Scattering. J Phys Chem B 104: 3405-3406.

Ghosh R, Bagchi B (2016). Enhanced density fluctuations in water-ethanol mixtures at low ethanol concentrations: Temperature dependent studies. arXiv.orgcond-mat > arXiv:1605.00386.

Ho M-W (2014). Large Supramolecular Water Clusters Caught on Camera – A Review. Water 6:1-12.

Horne RA (1972). Water and Aqueous Solutions: Structure, Thermodynamic and Transport Processes. Wiley-Interscience New York. Chapters 8-12.

Klein SD, Sandig A, Baumgartner S, Wolf U (2013). Differences in median ultraviolet light transmissions of serial homeopathic dilutions of Copper Sulfate, Hypericum perforatum, and Sulfur. Evid Based Complement Alternat Med., Vol 2013, Article ID 370609, 11 pages, 2013.

Konovalov AI (2013). The formation of nanosized molecular ensembles in highly dilute aqueous solutions. Herald Russ. Acad. Sci., 83: 513–519.

Konovalov AI, Ryzhkina IS (2014). Reviews: Formation of nanoassociates as a key to understanding of physicochemical and biological properties of highly dilute aqueous solutions. Russ Chem Bull Int Ed 63: 1-14.

Konovalov DA, Murtazina LI, Ryzhkina IS, Konovalov AI (2015). Effect of Weak Electromagnetic Fields on Self_Organization of Highly Diluted Solutions of Alkylated p_Sulfonatocalix[6]arene. Dokl Phys Chem 463: 154–157.

Konovalov AI, Ryzhkina IS (2016). Nanoassociates as a possible basic element of scientific foundation of homeopathy 1st Eurasian Congress on homeopathic medicine and 25th Congress of Asian Homeopathic Medical League, Moscow 14-15 October 2016.

Kozumi T, Kitagawa Y (2016). Water structure changes induced by ceramics can be detected by increased permeability through aquaporin. Biochem Biophys Rep 5: 353–358.

Lo S-Y (1996). Anomalous state of ice. Mod Phys Lett 10: 909-919.

Lo S-Y, Geng XU, Gann, D (2009). Evidence of stable-water-clusters at room temperature and normal pressure. Phys Lett A 373: 3872 -3876.

Lobyshev VI, Tomkevich MS, Petrushanko I Yu (2005). Experimental Study of Potentiated Aqueous Solutions.  Biophys 50: 464-469.

Lunev IV, Khamzin AA, Popova II, Ovchinnikov MN, Ryzhkina IS, Mishina OM, Kiseleva YuV, Academician Konovalov AI (2014). Dielectric Spectroscopy Study of Low_Concentration Aqueous Solutions of a Calix[4]resorcinarene Derivative. Dokl Phys Chem 455:56–59.

Mahata CR (2013). Dielectric dispersion studies of some potentised homeopathic medicines reveal structured vehicle. Homeopathy 102:262-267.

Miranda AR, Vannucci A, Pontuschka WM (2011). Impedance spectroscopy of water in comparison with high dilutions of lithium chloride. Mater Res Innov 15:302-309.

Mishina OA, Murtazina LI, Ryzhkina IS, Konovalov AI (2015). The relationship between self-organization, physicochemical properties, and biological activity of low-concentration solutions of p-aminobenzoic acid. Russ Chem Bull Int Ed 64:590-596.

Montagnier L, Aïssa J, Ferris S, Montagnier JL, Lavallee C (2009). Electromagnetic signals are produced by aqueous nanostructures derived from bacterial DNA sequences. Interdiscip Sci Comput Life Sci 1: 81–90.

Montagnier L, Aïssa J, Del Giudice E, Lavallee C, Tedeschi A, Vitiello G (2011). DNA waves and water. J Phys: Conf Ser 306: 012007.

Montagnier L, Del Giudice E, Aïssa J, Lavallee C, Motschwiller S, Capolupo A, Polcari A, Romano P, Tedeschi A, Vitiello G (2015). Transduction of DNA information through water and electromagnetic waves. Electromagn Biol Med 34: 106–112.

Mulliken RS, Ermler WC (1981). Polyatomic molecules, Academic Press, New York, Chapter II.

Nickolaenko AP and Hayakawa M (2002). Resonances in the Earth-ionosphere cavity. Kluwer Academic Publishers, Dordrecht-Boston-London.

Pal’mina NP, Mal’tseva EL, Kurnakova NV, Burlakova EB (1994). Effect of α-Tocopherol over a wide concentration range (10-17-10-2M) on Protein Kinase C. Biochimiya (Russian) 59: 193-200.

Pal’mina NP, Chasovskaya TE, Ryzhkina IS, Murtasina LI, Konovalov AI (2009). Water solutions of phenosan potassium salt: influence on biological membrane structure and conductivity. Dokl Biochem Biophys 429: 301–304.

Pang X-F (2014). Water: Molecular Structure and Properties (World Scientific Publishing, Singapore), Ch. 3.

Pershin SM, Bunkin AF, Grishin MYa , Davydov MA, Lednev VN, Palmina NP, Fedorov AN (2015).  Correlation of Optical Activity and Light Scattering in Ultra_Low_Concentrated Phenosan–Potassium Aqueous Solutions. Dokl Phys 60:114–117.

Pershin SM, Bunkin AF, Grishin MYa , Lednev VN, Fedorov AN, Palmina NP, (2016).  Bimodal Dependence of Light cattering/Fluctuations on the Concentration of Aqueous Solutions. Phys Wave Phenom 24:41–47.

Preparata G (1995). QED Coherence in Matter. World Scientific, Singapore, New Jersey, London, Hong Kong.

Pynzar EI, Bogdanova NG, Palmina NP (1995). Effects of Phorbol Esters on lipid peroxidataion in endoplasmic reticulum membranes from mouse liver. Membr and Cell Biol 9:279-288.

Quickenden TI,  Irvin JA (1980) The ultraviolet absorption spectrum of liquid water. J Chem Phys 72:4416-4428.

Rey L (2003). Thermoluminescence of ultra-high dilutions of lithium chloride and sodium chloride. Physica A 323: 67–74.

Robinson RA, Stokes RH (2002). Electrolyte Solutions (Dover Publications Inc., Mineola, New York. Chapters 1-3.

Ryzhkina IS, Murtazina LI, Sherman ED, Valitova YuN, Kataev EA, Konovalov AI (2010). Low-Concentration Aqueous Solutions of Macrocyclic Pyridine-Pyrrole Compound: Relationship between the Parameters, Physicochemical Properties, and Physiological Activity of Supramolecular Nanosized Associates. Dokl Phys Chem 443: 142-146.

Ryzhkina IS, Murtazina LI, Sherman ED, Pantyukova ME, Masagutova EM, Pavlova TP, Fridland SV, Academician Konovalov AI (2011a). Physicochemical Substantiation of the Hormetic Response of Biosystems for Wastewater Treatment to the Action of Solutions of N,N_Diphenylguanidinium Bis(hydroxymethyl)phosphinate. Dokl Phys Chem 438: 98–102.

Ryzhkina IS, Murtazina LI, Academician Konovalov AI (2011b). Action of the external electromagnetic Field is the condition of nanoassociate formation in highly diluted aqueous solutions. Dokl Phys Chem  440: 201–204.

Ryzhkina IS, Kiseleva YuV, Murtazina LI, Pal’mina NP, Belov VV, Mal’tseva EL, Sherman ED, Timosheva AP, Academician Konovalov AI (2011c). Effect of α-Tocopherol concentrations on the self-organization, physicochemical properties of solutions, and the structure of biological membranes. Dokl Phys Chem 438:109–113.

Ryzhkina IS, Kiseleva YuV, Murtazina LI, Academician Konovalov AI (2012a). Effect of ultralow concentrations and electromagnetic fields. Dokl Phys Chem 446: 153–157.

Ryzhkina IS, Kiseleva YuV, Timosheva AP, Safiullin RA, Kadirov MK, Valitova YuN, Academician Konovalov AI (2012b). Low-Concentration Aqueous Solutions of an Amphiphilic Calix[4]resorcinarene Derivative: Self-Organization, Physicochemical Properties, and Biological Activity under Common and Hypoelectromagnetic Conditions. Dokl Phys Chem 447: 193-199.

Ryzhkina IS, Murtazina LI, Masagutova EM, Mishina OA, Pavlova TP, Fridland SV, Academician Konovalov AI (2012c). Self-Organization of Sodium Chloride Solutions in the Absence and Presence of a Biologically Active Substance of Low Concentration under Common and Hypoelectromagnetic Conditions. Dokl Phys Chem 446:184-189.

Ryzhkina IS, Mishina OA, Timosheva AP, Segevaa SY, Kravchenko AN, Murtazina LI, Academician Konovalov AI (2013). Correlations between the self-organization, physicochemical properties and biological activity of Mebicar in dilute aqueous solutions. Mendeleev. Commun 23: 262-264.

Ryzhkina IS, Kiseleva YV, Murtazina LI, Mishina OA, Timosheva AP, Sergeeva SY, Baranov VV, Kravchenko AN, Konovalov AI (2015a). Self-organization and chirality in the high dilution solutions of glycoluril enantiomers with (R)- and (S)-methionine moieties. Mendeleev Commun 25:72–74.

Ryzhkina IS, Murtazina LI, Kiseleva Yu V, Konovalov AI (2015b). Self-Organization and Physicochemical Properties of Aqueous Solutions of the Antibodies to Interferon Gamma at Ultrahigh Dilution. Dokl PhysChemi 462: 110–114.

Ryzhkina IS, Kiseleva YuV, Mishina OA, Murtazina LI, Litvinov AI, Kadirov MK, Konovalov AI (2015c). Self-organization and properties of dilute aqueous solutions of cetyltrimethylammonium bromide in a range of physiologically important temperature. Russ Chem Bull Int Ed 64:579-589.

Samal S, Geckeler KE (2001). Unexpected solute aggregation in water on dilution. Chem Commun 2224–2225.

Sedlak M (2006). Large-scale supramolecular structure in solutions of low molar mass compounds and mixtures of liquids: Light scattering characterization. J Phys Chem B 110: 4329-4338; 4339-4345; 13976-13984.

Sedlak M, Rak D (2013). Large-scale inhomogeneities in solutions of low molar mass compounds and mixtures of liquids: Supramolecular structures or nanobubbles? J Phys Chem B 117: 2495-2504.

Segarra-Martí J, Coto PB, Rubio M, Roca-Sanjuán D, Merchán M (2013). Towards the understanding at the molecular level of the structured-water absorption and fluorescence spectra: A fingerprint of π-stacked water. Mol Phys 111(9-11):1308–1315.

Segarra-Martí J, Roca-Sanjuán D, Merchán M (2014). Can the hexagonal ice-like model render the spectroscopic fingerprints of structured water? Feedback from quantum-chemical computations. Entropy 16:4101-4120.

Upadhyay RP, Nayak C (2011). Homeopathy emerging as nanomedicine. Indian Journal of Homeopathy, 2012, Vol. 6, Issue: 3, Page: 31-38.

Vitiello G (2012). Fractals, Dissipation and Coherent States in Quantum Interaction. Eds. J R Busemeyer, F Dubois, A Lambert-Mogiliansky, M Melucci Vol.7620 pp.68-79, Springer Berlin Heidelberg.

Vitiello G (2014). On the Isomorphism between Dissipative Systems, Fractal Self-Similarity and Electrodynamics. Toward an Integrated Vision of Nature. Systems 2: 203-216.

Voeikov VL, Yablonskaya OI (2015). Stabilizing effects of hydrated fullerenes C60 in a wide range of concentrations on luciferase, alkaline phosphatase, and peroxidase in vitro. Electromagn Biol Med 34: 160–166.

Witt CM, Lüdtke R, Weisshuhn TE, Quint P, Willich SN (2006). The role of trace elements in homeopathic preparations and the influence of container material, storage duration, and potentisation. Forsch Komplementmed. 13:15-21.

Wolf U, Wolf M, Heusser P, Thurneysen A, Baumgartner S (2011). Homeopathic Preparations of Quartz, Sulfur and Copper Sulfate Assessed by UV-Spectroscopy. Evid Based Complement Alternat Med. Vol. 2011:692798. doi: 10.1093/ecam/nep036.

Yinnon CA, Yinnon TA (2009). Domains in aqueous solutions: theory and experimental evidence. Mod Phys Lett 23: 1959-1973.

Yinnon TA, Yinnon CA (2011). Electric dipole aggregates in very dilute polar liquids: Theory and experimental evidence. Int J Mod Phys B25: 3707-3743.

Yinnon TA, Yinnon CA (2012). Domains of solvated ions in aqueous solutions, their characteristics and impact on electric conductivity: theory and experimental evidence. Mod Phys Lett 26: 1150006-1 to 1150006-14.

Yinnon TA, Elia V (2013). Dynamics in perturbed very dilute aqueous solutions: theory and experimental evidence.  Int J Mod Phys B27:1350005-1 to 1350005-35.

Yinnon TA, Liu Z-Q (2015a). Domains Formation Mediated by Electromagnetic Fields in Very Dilute Aqueous Solutions: 1. Quantum Electrodynamic Aspects. Water 7:33-47.

Yinnon TA, Liu Z-Q (2015b). Domains Formation Mediated by Electromagnetic Fields in Very Dilute Aqueous Solutions: 2. Quantum Electrodynamic Analyses of Experimental Data on Strong Electrolyte Solutions. Water 7:48-69.

Yinnon TA, Liu Z-Q (2015c). Domains Formation Mediated by Electromagnetic Fields in Very Dilute Aqueous Solutions: 3. Quantum Electrodynamic Analyses of Experimental Data on Solutions of Weak Electrolytes and Non-electrolytes.  Water 7: 70-95.

Yinnon TA, Elia V, Napoli E, Germano R, Liu ZQ (2016). Water ordering induced by interfaces: an experimental and theoretical study. Water 7:96-128.

Zheltikov A, L’Huillier A, Krausz F (2007).  Nonlinear Optics, in Träger F, Ed. Springer Handbook of Lasers and Optics. Springer Berlin Heidelberg. pp157-248.

Zheng JM, Chin WC, Khijniak E, Khijniak E Jr., Pollack GH (2006). Surfaces and interfacial water: Evidence that hydrophilic surfaces have long-range impact. Adv Colloid Interface Sci 127: 19–27.

Zhigacheva IV, Evseenko LS, Burlakova EB, Fattakhov SG, Konovalov AI (2009). Effect of organophosphorus plant growth regulator on the electron transport in the mitochondrial respiratory chain. Dokl Biochem Biophys 427: 212-214.

Zubareva GM, Kargapolov AV, Yaguzhinskii LS (2003a). Fluctuations of pass band coefficients of water and salt solutions in the Infrared Region of the spectrum. Biofizika (Moscow) 48: 197-200.

Zubareva GM, Kargapolov AV, Laguzhinkii LS (2003b). Specific effect induced by subminute amounts of ascorbic acid on the fluctuation of transmission factor of water in the infrared spectral range. Dokl Biochem Biophys 388: 43-45

 


Table  1: List of abbreviations in alphabetic order, followed by Greek symbols abbreviations

Abbrev. Explanation

A/m – Amperes per meter

C – Concentration

Cx – X times centesimally diluted

– Critical concentration below which CDrot may form

– Transition concentration for CDplasma formation

– Transition concentration for IPDplasma formation

Cthr – Threshold concentration below which no domains are present in SDVSASES samples with concentrations above the Avogadro limit and screened by Permalloy.

– Coherence domain composed of coherent electronically excited water molecules

CDplasma – Coherence domain composed of few solvated solutes performing coherent plasma oscillation and numerous H20 molecules

CDrot – Coherence domains of ferroelectric ordered H20 molecules

DLS – Dynamic light scattering

DNA – Deoxyribonucleic acid

Ecoh – The energy a H2O gains by participating in the coherent oscillations of the molecules in a domain.

EDACDrot – Excited or broken CDrot piece, which has an electric dipole moment and therefore is denoted electric dipole aggregate (EDA)

EDAIPDplasma – Excited or broken IPDplasma piece, which has an electric dipole moment and therefore is denoted electric dipole aggregate (EDA)

EM – Electro magnetic

ELF – Extremely low frequency

EZ – Exclusion zone

eV – Electron Volt

H2O – Water molecule

H3O+ – Hydronium ion

Hz – Hertz

IPDplasma – In phase domains composed of few solvated solutes and numerous solvent molecules performing in phase plasma oscillation.

IR – Infra red

LiCl – Lithium Chloride

M – Molarity in mol per liter

m – Meter

mg – Milli gram

ml – Milli liter

NaCl – Sodium Chloride

NDS – Number of dilution steps

ng – Nano gram

nm – Nano meter

NMR – Nuclear magnetic resonance

OH– Hydroxide ion

Q – Heat of mixing of SDVSAS with acids or bases

QED – Quantum electro-dynamics

QFE – Quasi free electrons

SDVSAS – Serially diluted vigorously shaken aqueous solution

SDVSASLB – Serially diluted vigorously shaken aqueous solution kept on laboratory bench

– SDVSAS of anti-bodies kept at the laboratory bench

SDVSASP – Serially diluted vigorously shaken aqueous solution kept in Permalloy container

– SDVSAS of anti-bodies kept in Permalloy containers

Supra- – Agglomerate of 

Supra-CDrot – Agglomerate of CDrot

TGA – Thermogravimetric analyses

ULF – Ultra low frequency (500-3000 Hz)

UV – Ultra-violet

vc – Cyclotron frequency

vp – Frequency of plasma of quasi free electrons of 

σSurface tension

χElectrical conductivity in µS cm-1


 

Discussion with Reviewers

Reviewer A:

I am very curious if there is in the field any consideration on the possible role of gas bubbles inevitably included into the system during vigorous shaking procedure (necessary for manifestation of many of discussed properties of very diluted solutions). There are many of experimental evidences of existence of stable nanobubbles in water. If those gas inclusions are present and are filled with ambient air containing for example CO2 gas and water vapor (as in the theory of vesicles in a liquid presented by Prof. Pollack), both of those gases, CO2 and H2O, are so called greenhouse gases. Therefore, they absorb and re-emit infrared radiation and this radiation can stimulate formation of coherent domains of water. If we then cut-off external radiation we also cancel the effect of the gases. Could therefore gas bubbles be related to the effect of vigorous shaking on some properties of diluted solutions?

Yinnon TA:

Indeed, many researchers have considered the possibility that gas bubbles affect or even underlie the specific properties of SDVSAS. During preparation of SDVSAS, in particular when the vigorous shaking is carried out by succussions, not just nano-bubbles are created but the formation of gas bubbles is observable with the naked eye. Of course, the bubbles affect some physicochemical properties of the liquid. Your conjecture that gas bubbles enhance formation of coherent domains indeed is plausible. However, by analyzing extensive data, researchers have shown that the specific physicochemical and bioactive properties of SDVSAS are not attributable to gas bubbles alone. Below, I suffice with presenting only a few arguments which reject the option that gas bubbles underlie all the measured physicochemical and bioactive properties of SDVSAS.

i. As noted in the introduction, for aqueous solutions of many but not all compounds which are bioactive at physiological concentrations (about 10–2 to10–7 mol/liter), their serial dilutions beyond a transition concentration (Cthr), when combined with vigorous shaking after each dilution step, endow these with a specific bioactivity. Cthr depends on the solute. Solute specific bioactivity has been observed for SDVSAS diluted up to the Avogadro limit and beyond it. For such highly diluted SDVSAS, their bioactivity is correlated with their various physicochemical properties. Moreover, their bioactivity non-monotonically changes with the number of dilution steps. Sometimes, the bioactivity of SDVSAS at concentrations above Cthr differs from that at concentrations below Cthr. For example, Ryzhkina et al. (2013) have demonstrated that SDVSAS of Mebicar at 1×10-3 mol/liter exhibits an anxiolytic effect in mice, i.e., it provides a sedative effect combined with a decrease in exploratory activity while retaining the motion activity. However, at concentrations below 10-5 mol/liter, the liquid has a psychostimulating effect, i.e., it increases the motion and exploratory activities of the mice. The physicochemical properties of this liquid show that Cthr≈10-5 mol/liter (Ryzhkina et al., 2013) — in other words, on screening the liquid from ambient EM fields, at concentrations below this Cthr, its molecular associates disappear and its physicochemical properties become those of the solvent. At 1×10-7 mol/liter, the psychostimulating effect of the SDVSAS of Mebicar has a maximum. At concentrations of 1×10-10 and 1×10-14 mol/liter, the psychostimulating effect decreases. However, at 1×10-14 mol/liter, the motion activity increases.

On shielding the SDVSAS of Mebicar from ambient EM fields, at concentrations below Cthr≈10-5 mol/liter, Ryzhkina et al. (2013) observed: “the mean pH values of the solutions (pH 6) do not change, the Δε valuesi of the solutions approach zero; and α (the optical activity) does not reliably exceed the measurement error.” Moreover, these researchers observed that the screening led to the disappearance of the 10-7 – 10-4 m sized molecular associates. Accordingly, Ryzhkina et al. (2013) pointed out: “These findings serve as evidence that, in spite of the possibility of the influence of such contaminations as CO2 from the atmosphere or Ca2+ and silicate ions from glass, the true cause of the nonlinear concentration dependences of physicochemical properties is the formation and rearrangement of nanoassociates. If there are no nanoassociates, the properties of low-concentration Mebicar solutions correspond to the properties of distilled water.” The phenomena observed by Ryzhkina et al. (2013) are repeatable. They carried out their experiments in triplicate.

ii. The UV absorbance spectra of SDVSAS have the typical features of structured water (Elia et al., 2014c, 2015). Examples of structured waters are water adjacent to Nafion or other hydrophilic membranes [i.e., exclusion zone (EZ) water], aqueous solutions, water iteratively perturbed by Nafion membranes or by filtrations (Zheng et al., 2006; Chai et al., 2008; Elia et al., 2013a, 2014b-c). For aqueous solutions, i.e., those which were vigorously shaken as well as those which were not vigorously shaken, dynamic light scattering in combinations with other analyses revealed H2O ordered in sub-micron sized aggregates, which are not due to gas bubbles (Sedlak, 2006, 2013; Ryzhkina et al., 2012c; Konovalov, 2013; Ryzhkina et al. 2013; Konovalov and Ryzhkina, 2014). Examples for such aqueous solutions are alkali halide, sugar or alcohol solutions with concentrations in the range of 1 M to 0.1 M and the SDVSAS referred to in this paper.

SDVSAS, just as all the above mentioned structured waters, absorb in the 200 – 400 nm wavelength range. Typically, these waters have a peak in the 205 – 210 nm range. The peak is similar to the one observed for magnetized water (Pang, 2014). In addition, these waters have a broad peak in the 225 – 325 nm range. This broad peak has a maximum around 260 – 280 nm. A similar broad peak was first observed in water by Larzul et al. in 1965. Subsequent extensive research showed that pure water (gas phase H2O, bulk water, amorphous, hexagonal or cubic ice) does not absorb in the 200 – 400 nm range (Quickenden and Irvin, 1980; Mulliken and Ermler, 1981; Cabral do Couto and Chipman, 2012; Segarra-Martí, 2013).j Hence, the peaks in the spectra of water studied by Larzul et al. in 1965 have been attributed to some impurities.

For the 225 – 325 range peak, computations have recently been carried out with a high level, well established, quantum mechanical, ab initio derived theory, i.e., the Complete-Active-Space Self-Consistent-Field second-order perturbation theory (CASPT2). These computations indicate that the aggregates in structured waters are composed of excimers (Segarra-Martí et al., 2014). In particular, CASPT2 analyses of EZ water adjacent to Nafion membranes have show that these excimers form networks of multilayer honeycomb ice-like layers [see Fig. 1 and 6 in Segarra-Martí et al.(2014)]. The formula of the excimers is H38O20. The formula of the two monomers constituting the excimer is H19O10. The H2O forming the monomer are organized in two fused hexagons. In the H38O20 excimer, each of the two central (fused) H2O of one monomer attracts its opposite central (fused) H2O of the other monomer [See Fig. 6 in Segarra-Martí et al.( 2014)]. The attractive interaction is due to π-stacking. In the H38O20 networks, the π-stacked H2O resonate between their ground and an excited electronic state. About 10 percent of the H2O constituting the aggregates are simultaneously electronically excited.

The quantum mechanics CASPT2 results correspond to those obtained by the quantum dynamic model of structured waters, i.e., the quantum electro-dynamic model (QED). According to QED, the aggregates underlying the UV absorption in the 200-400 nm in structured waters are composed of coherent domains of the type denoted in this paper (Del Giudice et al., 2013; Elia et al., 2015 & 2017; Yinnon et al, 2015b&c, 2016). Within these domains, the H2O get electronically excited by ambient radiation with wavelength λ. These H2O deexcite by emitting photons which immediately get absorbed by H2O located within a distance λ. The subsequent repetitive excitation-deexcitation processes associate these H2O into an aggregate. About 10 percent of the H2O within a  simultaneously reside in their excited electronic state (Del Giudice et al., 2013).

A significant similarity between the quantum mechanical H38O20 and quantum dynamic models is their predictions concerning the following property of the H2O constituting the aggregates: About 10 percent of these H2O simultaneously reside in their electronic excited state.  One of the differences between the models is that the former predicts that the π-stacked H2O resonate between two of their electronic states (Segarra-Martí et al. 2014), while the latter predicts that all of the about 106 H2O composing a resonate between two of their electronic states (Del Giudice et al., 2013).

The aggregates in the structured waters are observable with fluorescence microscopy (Elia et al., 2013a, 2014c, 2017). Moreover, the aggregates are isolatable by evaporating in air drops of structured water or on lyophilizing (Lo et al., 2009; Ryzhkina et al., 2012b; Elia et al., 2013a, 2014b&c). The isolated aggregates are observable with atomic force microscopy and scanning electron microscopy (Lo et al., 2009; Ryzhkina et al., 2012b; Elia et al., 2013a, 2014b&c, 2017). The morphology of the isolated aggregates depends on the type of structured waters, their preparation techniques and concentration. State-of-the-art chemical analyses have indicated that these isolated aggregates mainly are composed of H2O (Elia and Napoli, 2010; Elia et al., 2017). On dissolving the isolated aggregates in Milli-Q water, the resulting liquid has physicochemical properties resembling those of the structured water from which it was prepared (Elia et al., 2014c, 2017). For example, its UV (absorption, fluorescence and circular dichroism) spectra have similar properties, e.g., the typical absorption peaks in the 200 – 400 nm range specified above. The fact that the chemical analyses have shown that the isolated aggregates are composed of H2O, implies that the aggregates are not  a kind of isolated bubbles. Of course, in the liquid phase, air bubbles may adhere to the aggregates.

Reviewer A:

Yet another notion, when vigorously shaking my solutions of ionic liquids (not very diluted and not related to the discussed field of investigation) I found it virtually impossible to avoid contamination by airborne fungi. I have been using sterile containers and filtered my solutions through 0.22 µm filters, yet after some time I could always observe fungi growth in solutions where they had organic feeding (ionic liquid) material. Is it possible that in case of non-sterile air mixed with dilute vigorously shaken samples we introduce fungi spores that can produce some artifacts in the analyzed samples?

Yinnon TA:

Paragraph ii of the Discussion with Reviewers Section partly is also relevant to this question. However, I would like to add the following remarks:

⚬ The group of Prof. Elia carried out many carefully controlled experiments to exclude the possibility that the differences between SDVSAS and aqueous solutions of equal chemical composition but prepared without serial dilutions and vigorously shaken are attributable to contaminants of any kind (Elia and Niccoli, 2004a&b; Ciavatta et al., 2008; Elia and Napoli, 2010).

⚬ Highly controlled, blinded and randomized experiments of UV radiation transmission by SDVSAS were carried out by Wolf et al. (2011) and Klein et al. (2013). Statistical analyses of the data have shown that the UV radiation’s transmission of these aqueous SDVSAS significantly differed from that of the controls. The controls were vigorously shaken solvents. These solvents were the various types of purified waters or the 99% water and 1% ethanol mixture with which the SDVSAS were prepared. The purified waters included distilled water, Quartz distilled water and de-ionized water. Part of the experiments were carried out in a metal-free class 100 High Efficiency Particulate Air (HEPA) filtered clean room. Clean flow boxes had class 5. These experiments showed that the UV radiation’s transmission by SDVSAS of copper sulfate (CuSO4),  or aqueous SDVSAS of hypericum is lower than that of the control. However, for SDVSAS of sublimed sulfur (S8), it is higher than the control.

Reviewer B:

How does the author correlates her description of SDVSAS to the recent finding of V. Elia and collaborators concerning the supramolecular aggregates of water molecules remaining solid at ordinary pressure and room temperature both in homeopathic solution and also in other pure water physical treatments (iterative filtration, iterative exposure to Nafion…) presented in Elia et al. (2013a, 2014b, 2015, 2017) ?”

Yinnon TA:

In paragraph ii of the Discussion with Reviewers Section, I already mentioned that the group of V. Elia has shown that iteratively perturbing polar liquids (e.g., water, alcohol or their solutions) may lead to stabilization of 10-7 – 10-4 m sized supramolecular aggregates, which mainly are composed of solvent molecules. Such aggregates may stabilize when such liquids are iteratively perturbed by:

□ serial filtrations (Elia et al., 2014b, 2015);

□ repeated exposure to Nafion (Elia et al., 2013a&b, 2014a, 2015, 2017).

□ repeated shaking after each dilution step of serially diluted liquids, i.e., the method used for preparing homeopathic liquids and SDVSAS (Lo et al., 2009; Upadhyay and Nayak, 2011; Konovalov, 2013; Elia et al., 2014c, 2015).

These aggregates are isolatable by evaporating-in-air drops of the perturbed liquid or by lyophilizing it. Such evaporation techniques leave a solid residue. For water perturbed by serial filtrations or repeated exposure to Nafion, the residue is composed of H2O and contains minuscule amounts of impurities (organic contaminants and compounds released by the container or the Nafion).

Analyses of the physicochemical properties of the residues have revealed that their aggregates have characteristics of QED coherent domains (Elia et al., 2015, 2017; Yinnon et al., 2016). In particular, the thermogravimetric analyses of the residue of water perturbed by repeated exposure to Nafion have revealed some of these characteristics. The thermogravimetric graph shows that the residue contains two types of H2O aggregates. One type has the thermodynamic properties of CDrot and the other type has the thermodynamic properties of  (Yinnon et al., 2016). For example, the thermogravimetric graph reflects the critical temperatures below which these domains form. These measured temperatures are in good agreement with those predicted by classical electro-dynamic and quantum electro-dynamic (QED) models of aqueous systems.

The currently available analyses of water iteratively perturbed by Nafion or serial filtrations indicate that their aggregates, i.e., their CDrot and , are clumps of EZ water (Elia et al., 2015, 2017; Yinnon et al., 2016). These aggregates became stabilized by Nafion or the filter, i.e., constitute the interfacial EZ water. The iterative perturbations tore clumps of these aggregates from the  interfacial zone and dispersed these in the liquid.

My QED model of SDVSAS points out that serially diluting polar liquids below a transition concentration ()  may create a QED domain, which I denoted IPDplasma. In the paper by Yinnon and Yinnon (2012), the possibility for formation of IPDplasma was derived for the first time. In that paper, experimental data in support of this possibility was cited. (After the paper was published, the late Prof. Del Giudice wrote me that he had been one of the reviewers.) Also the current paper’s analyses of experimental data, as well as the analyses by Yinnon and Liu (2015b&c), show that the characteristics of IPDplasma conform to those observed for aggregates present in aqueous solutions. The effects of vigorous shaking of a solution on their IPDplasma are first analyzed in the paper by Yinnon and Yinnon (2011). The central finding of this paper is that shaking may cause creation of aggregates with an electric dipole moment (EDA). In the Yinnon and Yinnon (2011) paper, it also is pointed out that on further diluting below a critical concentration  the EDA or solutes with a sizable electric dipole moment may stabilize CDrot and . As discussed in the current paper and my previous ones on SDVSAS, the CDrot and underlie these liquids’ physicochemical properties. In other words, it was the theoretical discovery of IPDplasma and EDA, in combination of my identification of the extensive experimental data in the literature confirming the presence of aggregates with characteristics of IPDplasma, , EDA, CDrot and which enabled me to provide consistent explanations for the many observed physicochemical properties of serially diluted vigorously shaken polar liquids. Here, I emphasized the importance of the theoretical identification of IPDplasma, and EDA for leading the way to understand SDVSAS. Just claiming that the QED model of aqueous systems is more accurate than the customarily electrostatic models is not sufficient for explaining the properties of SDVSAS. The effects of vigorous shaking on aggregates resulting from QED interactions have to be carefully assessed.

The previous paragraphs accentuate that the current available analyses all show that the CDrot and  QED domains may become stabilized by iteratively perturbing water by serial filtrations or by repeated exposure to Nafion, as well as by repeated shaking after each dilution step of serially diluted liquids. Stabilization of such domains implies breaking of the symmetry of the liquid. The group of G. Vitiello has extensively studied such symmetry breaking. Examples are the following excerpts in the paper “Classical Limit and Spontaneous Breakdown of Symmetry as an Environment Effect of Quantum Field Theory” by Celghini, Graziano and Vitiello Phys Lett A 145:1-6 (1990):

→ “Small reaction fields arising from the system-environment coupling are shown to drive the system into an asymmetric ground state in the group contraction limit. This limit acts as the theory of classical limit.”

→ We believe it is therefore worthwhile to move the discussion on small reaction fields from the quantum mechanics setting to QFTk which is the natural framework for phase transition and spontaneous breakdown of symmetry, and …”

→ “Every very tiny perturbation arising from the system-environment interaction (reaction field) may trigger the breakdown of symmetry thus producing long range correlations in the system.”

→ “The system order parameter does not depend on the details of the environment state.”

→ “Finite volume effect set a threshold above which the intensity of the reaction field must be in order to have symmetry breakdown.”

→ “The discussion presented in this paper shows that in QFT small reaction fields from the environment may play a crucial role in triggering the breakdown of continuous symmetry with the formation of ordered patterns in the system.”

The above implies that the effects of iterative perturbations of water, polar liquids and their solutions, as analyzed by the group of V. Elia, A. Konovalov and others in the years 1999 – 2017 indeed conform to those predicted by the group of G. Vitiello in 1990.

i Δε is the difference between the dielectric constant of this SDVSAS and that of the pure solvent, i.e., twice distilled water.

j For gas phase H2O, its lowest-energy band of the electronic spectrum covers the 151 – 182 nm range, with a maximum at 168 nm (Segarra-Martí et al. 2013; Cabral do Couto and Chipman, 2012). In pure bulk liquid, this band is broader and blue-shifted (Quickenden and Irvin,1980; Segarra-Martí et al. 2013; Cabral do Couto and Chipman, 2012). Its maximum is at 151 nm and its absorbance falls off monotonically by about ten orders of magnitude in the 151 – 400 nm range [it is only about 0.0001 cm-1 at 320 nm — see (Quickenden and Irvin,1980)]. Amorphous, hexagonal or cubic ice have absorbance features similar to those detailed in the previous sentence (Segarra-Martí et al. 2013; Cabral do Couto and Chipman, 2012).

k QFT is the abbreviation of Quantum Field Theory. The QED model of aqueous systems and other polar liquids is derived within the context of QFT.

 
Close Menu