biophysical interactions: their paramount importance for life

15
Romanian Reports in Physics, Vol. 65, No. 3, P. 1063–1077, 2013 Dedicated to Professor Valentin I. Vlad’s 70 th Anniversary BIOPHYSICAL INTERACTIONS: THEIR PARAMOUNT IMPORTANCE FOR LIFE A.I. POPESCU 1 , D.M. GĂZDARU 1 , C.G. CHILOM 1 , M. BACALUM 1,2 1 University of Bucharest, Department of Electricity, Solid State and Biophysics, Faculty of Physics, 405 Atomistilor, Măgurele-Ilfov, ROMANIA 2 “Horia Hulubei” National Institute of Physics and Nuclear Engineering, Department of Life and Environmental Physics, Măgurele, PO Box MG-6, 077125, ROMANIA Corresponding author: [email protected] Received June 12, 2013 Abstract. Life processes are manifesting properly only by correlated and multiple biomolecular interactions taking permanently place in all the cells. The sine qua non molecular interactions for life processes are the specific interactions, that is, those conducting to preferential associations between partners. These interactions are the result of many synergic factors. The common specific interactions are the enzyme- substrate and the protein-ligand interactions. In the case of enzymes, the natural ligands are the specific substrates bound to enzyme active sites before they are split into products. Quantitatively, the specific interactions are characterized by the association (affinity) constant, a high affinity meaning a great value for this constant. The specific interactions are also ubicuitary both at supramolecular and cellular levels facilitating a coherent interplay of all the living matter components. Indeed, a large variety of subtle biological phenomena taking place at cellular level, are involving cell-to-cell interactions, more specifically, membrane-to-membrane interactions of the apposing cells. The cell-to-cell interactions are mainly determined by the composition and spatial structure of their membrane and their membrane peripheries (i.e. glycocalyx) and also by the particular composition of the liquid micro-environment. Some specific interactions both at molecular (first part) and cellular (second part) level are shortly reviewed and commented. Key words: biomolecular interactions, steric and electrostatic complementarity, protein- ligand, affinity constant, cell-to-cell interactions, electrostatic and electrodynamic interactions, hydration forces. 1. INTRODUCTION A fundamental question addressed by biophysics is the following: which are the simplest processes considered as “elementary” ones sustaining the Life processes? Alike any other natural phenomena Life is maintained by myriad of

Upload: others

Post on 09-Feb-2022

0 views

Category:

Documents


0 download

TRANSCRIPT

Romanian Reports in Physics, Vol. 65, No. 3, P. 1063–1077, 2013

Dedicated to Professor Valentin I. Vlad’s 70th Anniversary

BIOPHYSICAL INTERACTIONS: THEIR PARAMOUNT IMPORTANCE FOR LIFE

A.I. POPESCU1, D.M. GĂZDARU1, C.G. CHILOM1, M. BACALUM1,2 1 University of Bucharest, Department of Electricity, Solid State and Biophysics, Faculty of Physics,

405 Atomistilor, Măgurele-Ilfov, ROMANIA 2 “Horia Hulubei” National Institute of Physics and Nuclear Engineering, Department of Life

and Environmental Physics, Măgurele, PO Box MG-6, 077125, ROMANIA Corresponding author: [email protected]

Received June 12, 2013

Abstract. Life processes are manifesting properly only by correlated and multiple biomolecular interactions taking permanently place in all the cells. The sine qua non molecular interactions for life processes are the specific interactions, that is, those conducting to preferential associations between partners. These interactions are the result of many synergic factors. The common specific interactions are the enzyme-substrate and the protein-ligand interactions. In the case of enzymes, the natural ligands are the specific substrates bound to enzyme active sites before they are split into products. Quantitatively, the specific interactions are characterized by the association (affinity) constant, a high affinity meaning a great value for this constant. The specific interactions are also ubicuitary both at supramolecular and cellular levels facilitating a coherent interplay of all the living matter components. Indeed, a large variety of subtle biological phenomena taking place at cellular level, are involving cell-to-cell interactions, more specifically, membrane-to-membrane interactions of the apposing cells. The cell-to-cell interactions are mainly determined by the composition and spatial structure of their membrane and their membrane peripheries (i.e. glycocalyx) and also by the particular composition of the liquid micro-environment. Some specific interactions both at molecular (first part) and cellular (second part) level are shortly reviewed and commented.

Key words: biomolecular interactions, steric and electrostatic complementarity, protein-ligand, affinity constant, cell-to-cell interactions, electrostatic and electrodynamic interactions, hydration forces.

1. INTRODUCTION

A fundamental question addressed by biophysics is the following: which are the simplest processes considered as “elementary” ones sustaining the Life processes? Alike any other natural phenomena Life is maintained by myriad of

A.I. Popescu et al. 2 1064

interactions between a large variety of particles: ions, small molecules, mesomolecules, macromolecules, supramolecular complexes, cells, etc. Therefore one could say that the “elementary processes” sustaining Life are the interactions taking place within an intracellular multi compartmented structure embedded into an aqueous medium. But, unlike the processes taking place/occurring in the inert matter, Life is especially sustained and driven by specific interactions (i.e., selective interactions) favouring the preferential associations between the partners.

2. MOLECULAR INTERACTIONS

The molecular interactions are due to stochastic collisions between molecules, the complementary structure of these favouring or not the formation of high order molecular complexes. But even in the particular case of a perfect complementarity of the two reacting agents, a single collision has a very small probability to form a complex, due to the fact that the two colliding partners must fulfill favourable anisotropy conditions (e.g., collision direction and angle). But, the collision repetition of hundreds of millions times a second as well as the huge number of participating molecules determines that this very small probable possibility of complex formation to become an effective one in real time.

Therefore the completion of an interaction is due to the enormous number of molecular collisions promoted by thermal agitation and essentially accomplished by affinity of the partners determined by their steric and electrostatic complementarity.

2.1. THERMAL AGITATION AS A PROMOTER OF MOLECULAR INTERACTIONS

A deep insight into the biological phenomena is indicating that Life is tributary to thermal agitation of the ions, molecules, macromolecules, supramacromolecular structures and even of the cells. The thermal agitation is favouring the collisions (i.e., interactions) between different particles of the biosystems being also responsible for the macromolecular “breathing” absolutely necessary for their operation. However, it is important to note the existence of a narrow temperature interval (tmin, tmax), especially for the homeothermic organisms, inside of which the thermal agitation is optimal. Bellow tmin, the molecular collisions are less frequent and weak and, consequently, the interactions are slown down inhibiting thus the physiological processes or even blocking them, if t << tmin. On the contrary, beyond tmax, the collisions are intensified (i.e., they become more frequent and more powerfull) leading to the physiological processes acceleration. For t >> tmax, the collisions are so violent that the formed molecular complexes are disintegrated and/or the macromolecules are denatured.

3 Biophysical interactions: their paramount importance for life 1065

Integrity of the essential macromolecules (i.e., proteins, nucleic acids and glycans) as well as of their complexes is assured only by a moderate thermal agitation in the narrow temperature interval (tmin, tmax). For the homeothermic organisms, this narrow thermal interval is smaller than few degrees. In the case of the human beings, it encompasses maximum one degree: (36.5-37.5) ºC.

2.2. PARAMETERS DESCRIBING MOLECULAR INTERACTIONS

The cellular functions of most proteins are controlled by smaller molecules called ligands that reversibly bind to proteins stimulating or inhibiting their activity.

For the sake of simplicity, we shall analyze a generic reaction between two partners: the macromolecule, M (e.g., a protein, a nucleic acid) and the ligand, L (a small molecule, another protein, etc.). In the particular case of 1:1 stoichiometry, one can write the reversible reaction:

D

R

k

kM L ML→+ ← , (1)

where ML is the formed complex, kD and kR are the so called rate constants of direct reaction (association) and reverse reaction (dissociation) respectively.

This reaction is characterized by the affinity constant, KA:

[ ] /[ ][ ]DA

R

kK ML M L

k= = . (2)

If the chemical equilibrium is shifted to the formation of the complex ML (i.e., the concentration of the complex is high), one means that the two partners have a high reciprocal affinity and conversely. Quantitatively, a high affinity means a great value for KA. Depending on the specific properties of M and L, the affinity constant for specific interactions, can vary over many orders of magnitude (usually, from ~ 104

to 1010 M-1). For example, guanilate monophosphate kinase

binds the ligand, guanosine monophosphate, with a KA of ~ 104 M-1[1], human and bovine serum albumins binds eosin Y with a KA of ~ 106 M-1 [2], the Xeroderma pigmentosum group C protein binds to human centrin 2, with a KA of ~ 108 M-1 [3]. A higher specific interaction between IgG antibodies and their antigens is characterized by a higher affinity constant: KA ~ 109

M-1 [4]. It is interesting to note

that in streptavidin-biotin interaction the affinity constant is very high (KA = 1014 M-1), being one of the strongest non-covalent interactions in Nature [5].

Phenomenological approaches [6–8] of molecular interactions are performed, among other techniques, by isothermal titration calorimetry (ITC) who uses the thermodynamic state functions: enthalpy, H, entropy, S, and Gibbs free energy, G. From the first thermodynamic principle, one can deduce that, at constant pressure,

A.I. Popescu et al. 4 1066

the enthalpy variation is equal to the exchanged heat, Q (∆H = Q) . It means that the heat exchange can be measured as the enthalpy variation, ∆H. The second thermodynamics principle imposes, for the spontaneous processes, the decrease of Gibbs free energy (∆G < 0), and the increase of entropy (∆S > 0).

The biomolecular interactions are either energetic driven (endergonic), if ∆G > 0, taking place in vivo, or spontaneous (exergonic) if ∆G < 0. Only the exergonic interactions are spontaneous, while the endergonic ones must be driven by coupling with the exergonic ones. Therefore, in vitro, only the case of exergonic reactions is encountered. At constant pressure, the variation of ∆G is given by the known relation:

G H T S∆ = ∆ − ∆ , (3)

where G∆ , H∆ and S∆ represent the variations of G, H and S, and T is the absolute temperature.

At molecular level, Eq. (3) is reflecting two processes: the trend to decrease the energy (the binding formation: ∆H < 0) and the thermal movement (the breaking of different bindings: ∆S > 0).

The Gibbs free energy variation is related to the constant affinity by the relation:

ln AG RT K∆ = − , (4)

where R is the ideal gas constant. From the last two relations one can calculate the variation of the entropy, too:

( ln ) /AS H RT K T∆ = ∆ + . (5)

Eqs. (4) and (5) allow the determination of both ∆G and ∆S from ITC data from which one appreciate if the reaction is entropically, enthalpically (or both) driven.

It is important to emphasize that the affinity constant is a macroscopic property of binding partner, representing an averaged behaviour of a very large number of “elementary” events (i.e. numerous microscopic non-covalent molecular interactions).

The strength of an interaction depends on the similarity and complementarity of the physico-chemical properties of atoms from “surfaces” of partners (e.g., protein and ligand). The molecular similarity is more precisely described as atom pair matching, involved in weaker interactions and electrostatic fingerprint charge matching, involved in stronger interactions [9].

2.3. SPECIFIC MOLECULAR INTERACTIONS

As it already said the myriad of biomolecular interactions are facilitated by thermal agitations that permit the particles to collide millions or billions times a second, the thermal background being a sine qua non condition for life manifestation.

5 Biophysical interactions: their paramount importance for life 1067

The most important molecular interactions for life processes are the specific interactions, the result of shape and charge complementarity of the two partner “surface” structures.

To emphasize these ideas, we shall exemplify the specific activation of a small protein, calmodulin (CaM), by four calcium ions which are binding in specific sites (Fig. 1):

Fig. 1 – An example of cascade activation by specific interactions: 4 cations are activating calmodulin (CaM) by their binding to 4 CaM specific sites; the activated calmodulin (CaM*) in its turn, is specifically interacting with the enzyme, E, which becomes activated, exposing its active centre (AC) to the substrate, S, which will be

split in the product, P.

CaM + 4 Ca2+ ↔ CaM*. (6)

The activated calmodulin, CaM*, is activating tens of enzymes, E, otherwise remaining in dormant inactive states:

CaM* + E ↔ CaM + E*. (7)

In fact, CaM* is specifically interacting with a polypeptidic portions of the regulatory domains of the enzymes [10].

A.I. Popescu et al. 6 1068

Finally, the activated enzyme, E*, is able to specifically interact, by its active centre, with its specific substrate, S, forming an activated complex, E*S≠, in which the generated mechano-electrical tensions will split the substrate, resulting the product, P:

E*+S E*S E * +PD P

R

k k

k≠→ →← , (8)

where, kD and kR are the rate constants of the direct and reverse reaction and kP the rate constant of the E*S≠ dissociation into a recyclable enzyme and the product, P.

The specific interactions are the result of many synergic factors: a) steric shape complementarity of the partners; b) complementarity of electrical charges of the two partner surfaces; c) extensive network of hydrogen bonds between the partners; d) the numerous van der Waals interactions; e) the hydrophobic interactions; f) and, in certain cases, the rigidization of a flexible loop which closes over one of the bound partner [11].

Simple binding has to do with similarity so that, the higher similarity is between the interacting partners, the more specific the recognition will be.

There is evident that many proteins dynamically associate with other similar or different ones in order to perform certain biological functions [12]. For instance, transient protein-protein interactions play an essential role in signal transduction pathways [13] as is the case of a large class of membrane proteins called G-protein-coupled receptors. These receptors are involved in cellular responses to external signaling molecules (e.g., hormones, neurotransmitters) and operate either as homo- or hetero-oligomers [14, 15].

2.4. PERSPECTIVES

We shall mention only three applications of specific molecular interactions: a) molecular recognition; b) antibody-antigen interactions, and c) rational drug design.

a) Molecular recognition is a complex and fundamental process, essential for Life. However, ligand-transmembranar receptor recognition is very difficult to be characterized because one of the partners is embedded within cell membranes. But, with the advent of ITC one can approach the study of such a process [3, 16, 17].

b) Antibody-antigen interactions are characterized by large values of affinity constants (KA >109 M-1). A thermodynamic analysis provides a complete source of information used to interpret the multitude of molecular forces involved in the antibody-antigen coupling. It is expected that the future researches will be focused on design of more potent antibodies for use in cancer immunotherapy.

c) Rational drug design. ITC will play a leading role in molecular characterization of binding mechanisms in order to apply it in intelligent drug design [18]. Proteins are targets for different existing drugs and for yet unknown

7 Biophysical interactions: their paramount importance for life 1069

ones waiting to be synthetised. Reactions involving the binding of drugs to the protein active sites can be easily characterized by ITC [19,20]. To find an optimal drug for blocking a macromolecule (protein or nucleic acid) one needs to know the structure of its natural agonists/antagonists as well as the surface electrical charges and topology of the binding sites on that macromolecule.

3. CELLULAR INTERACTIONS

Generally, one can speak of permanent cellular interactions (e.g., interactions sustaining the tissue cells) and transient cellular interactions (e.g., interactions of free immunocompetent cells or interactions of bacteria with the tissue cells).

Cellular interactions are determinant in a huge variety of biological processes: specific associations (e.g., segregation, contact inhibition, differentiation, cellular immunity, cell fusion, etc.). In all these complex phenomena, the major role is played by the external surface of the cell membranes. Cell- -cell interaction is the resultant of many types of forces and is modulated by a large diversity of microenvironment factors [21, 22]. It is necessary also, to mention the presence of specific receptors on the cell membranes and the influence of different macromolecules in the cell suspending medium on cell interactions. Due to the fact that the biochemical composition and the architecture of the cell surfaces are different for various types of cells, the behaviours of cells in interaction with other cells are also various and specific. However, in spite of this large diversity, one can notice common characteristics of this phenomenon, irrespective of the cellular species, conferring a general feature to this process [23].

Until now, there is not a single theoretical approach able to satisfactorily explain all the aspects of the cellular behaviour during and after interactions, because different authors are operating with more or less simplified models of the real cells.

3.1. MODELS OF CELL-TO-CELL INTERACTIONS

As compared to the largest macromolecule, the cells are huge bodies composed of billions and billions of molecules and macromolecules. Therefore the mutual interaction between two cells is much more difficult to be approached and treated from biophysical point of view. For this reason, a simpler but approximate way is to model the cell as a whole, taking into account only its external coat (i.e. its glycocalix). In essence, the cellular interactions are the result of the simultan manifestation of manifold interactions, both specific and nonspecific, between the molecules/macromolecules composing the different cell coats.

A.I. Popescu et al. 8 1070

The first cell models were those of rigid non-conducting spheric colloidal particles which stimulated the implementation of DLVO theory [24, 25] in understanding cell-to-cell interactions. These models represent only a crude approximation of the reality due to their incapacity to explain the stability of cell population at very short distance of separation, where the expression of interaction energy has a singular point (vide infra). Some physical models mimicking the interactions between cells and the influence of adsorbed ions on the interactions, in the guise of hydration forces [26], were the mica plates [27], immersed into electrolyte solutions. More realistic models, suitable for studying the interactions of the cells without glycocalyx, but still too simple, as compared to the cell membranes, are the lipid bilayers [28–29]. But the best suited models were the theoretical ones, in which both intrinsic (integral) and extrinsic proteins are present in the membranes provided with a glycocalyx of a known composition [30].

3.2. NONSPECIFIC REPULSIVE ELECTROSTATIC INTERACTIONS

In physiological conditions, all the cells carry a net negative charge at their surfaces where there are also positive charges, but the negative ones are in excess. The origin of electrical charges of intrinsic nature is mainly due the dissociation of the chemical groups of the membrane surfaces and of glycocalyx, generating both negative (e.g., carboxyl and sulphate groups) and positive (e.g., amino groups) charges. The main contribution to the negative charge of the cell periphery is due to glycophorins which carry approximately 90 % of the glycocalyx sialic acid which possesses carboxyl groups and to proteoglycans which contain one or more polysaccharides contributing with many negative charges to cell surfaces [25]. Another contribution to the cell electrical charges is of extrinsic origin due to differential adsorption of anions and cations onto the cell surfaces. It results that also by this mechanism, the negative charges on the cell surfaces is exceeding the positive ones, the anions being preferentially adsorbed. The charges exposed on the cell peripheries will generate around the cells electric double layers, the electrostatic interaction between the cells being, in fact, the interactions between the apposing electric double layers. There are many mathematical expressions of the electrostatic energy of interactions, for different cells and interacting conditions [30], an example of them describing the electrostatic repulsion energy, GESR (per unit area) between two identical cells with small surface potential, Ψ0, and with a glycocalyx of thickness, δ, embedded in a uni-univalent electrolyte, at the distance, r, between their surfaces [31]:

9 Biophysical interactions: their paramount importance for life 1071

2 20 0( ) 16 exp(2 ) exp ) /( ) 0,ESRG r n e r kT≈ Ψ δ − κ κ > (9)

where 0n is the counterion concentration, e the elementary charge, k – the Boltzmann constant, κ – the Debye-Hückel parameter, and T – the absolute temperature.

3.3. NONSPECIFIC ATTRACTIVE ELECTRODYNAMIC (VAN DER WAALS) INTERACTIONS

A simplified mathematical formula for van der Waals energy of interaction, GVDW, per unit area (considering the cell surfaces as semi-infinite planar medium) is given by [30, 31]:

2

( ) 0,12VDW

AG rr

= − <π

(10)

where A is the non-retarded Hamaker constant (10-22 – 10-19 J). As it results from formulae (9) and (10) the electrodynamic (van der Waals) forces are attractive, while electrostatic forces are, repulsive. In the DLVO theory, only these two terms were taken into account and cannot properly explain the interactions taking place at a very short distance of separation. Indeed, according to the expression (10), for a very small distance between the interacting particles, the van der Waals attraction increases indefinitely. Owing to this indefinite increasing of the interaction, the cells would attract more and more strongly as they are approaching each other and, finally, the fusion and collapse would be inevitable. However, this is not the case, because normally, the interacting cells are able to establish a true contact (e.g., in a tissue) preserving more or less their global integrity Therefore the interaction of the cells at very short distance of separations clearly must imply the interplay of very strong short-range repulsive forces which must counterbalance the electrodynamic attractive ones.

3.4. NONSPECIFIC SHORT RANGE REPULSIVE FORCES

All the researchers agree with the existence of the short range repulsive forces, that must counterbalance the electrodynamic ones, but there are still controversies about the nature of these forces. Some of researchers are stressing the importance of hydration forces, others are considering important the Lewis acid-base interactions that were also considered as hydration pressure [32]. While these types of forces are important in the case of rigid membrane (i.e., without in-plane and out-of-plane movements), in the case of flexible ones, in which the thermal fluctuations are present, such repulsive forces are attributed to so called entropic forces due to the steric repulsion of the hydrated lipids.

A.I. Popescu et al. 10 1072

3.4.1. Hydration forces

When the interacting cells are going to establish a firm contact (i.e., at a very short distance) a repulsive force is experienced due to the steric hindrance [32] of the intrinsic or grafted extrinsic [33] macromolecules in the coats of the two cells The repulsive interactions are the result of two opposite phenomena: the osmotic pressure tending to retain water molecules into intercellular space and the steric compression, tending to squeeze out water from this space. Steric forces will realize a steep repulsive barrier, stopping the further approach of the two cells and consequently, preventing the cell fusion (i.e., without intervention of fusogenic agents). If we are ignoring the steric forces, the cells experience very strong repulsive forces [28, 30, 35, 36] at very short distance, preventing the cell fusion. One considers that these ”hydration” repulsive forces are due to the clustering of water molecules onto bilayer/membrane lipids, caused by the strong dipole moments of water molecules, and also to the hydrated ions adsorbed onto the cell surfaces. Although these forces were early postulated, their experimental evidence was obtained latter [37–41]. These forces are explaining the stability of different bacterial cells [42], the pressure the tissues withstand from compression forces, and the role of corneal barrier mucus against contamination with hydrophilic bacteria [43]. The space dependence of the “hydration” energy is given by an exponential function [28, 38, 41]:

0( ) exp( / ),HRFG r G r= − λ (11)

where λ is the space decay constant (usually, in the range of 0.1 to 0.3 nm) and 0G is the hydration energy at “zero separation” (in fact, equal to Born repulsion distance).

It can be deduced that water molecules which cover the lipid bilayers and by extension, the cell membranes, generate strong repulsive forces that counteract the attractive electrodynamic forces, at very short distance of separation [28, 35, 38, 41].

3.4.2. Entropic forces

The lipid bilayers, the structural framework of the membranes, are not at all rigid structure, out-of-plane thermal motions of its components being permanently manifested. These motions must have consequences on the cell-to-cell interactions. The studies on both flexible and rigid bilayers/membranes could delineate the partial contributions of the “hydration” forces and of the “entropic” forces due to the thermal fluctuations [26]. The measurement of forces between the membranes shows an abrupt decrease of membrane hydration, in the case of rigidified bilayers. Therefore, the repulsive forces seems not to be given only to the hydration, but also to thermal fluctuations of the bilayer lipids.

11 Biophysical interactions: their paramount importance for life 1073

The conclusion is that the hydration forces are acting at very short distances (no further than about 0.03 nm), while at short distance (no longer than 0.5 nm) the repulsive forces are due to steric interactions of the hydrated lipid headgroups and, for distances longer than 0.5 nm, the dominant interactions are those due to the “entropic” forces provoked by thermal out-of-plane motions.

The hydration short range repulsive forces are those that amend the DLVO theory, adding to the expression of total energy, the hydration energy, GHRF (r):

( ) ( ) ( ) ( ),CT ESR VDW HRFG r G r G r G r= + + (12)

explaining thus, the cell stability when they are realizing a firm contact at a distance corresponding to the minimum, m1 (Fig. 2). This minimum is situated at a distance equal to the sum between Born repulsion barrier (of about 1.4 Å, which prevent atoms to approach each other) and double length, 2δ, of cell glycocalyx [32].

GT BR BR

BR GCT

GHRF

GVDW < 0

GESR > 0

GHRF > 0

rr

r

r2

r2

m1

m1m2

m2

(Born Repulsion)GT

d

Fig. 2 – Expected qualitative representation of interaction energies as a function of distance, r, between two identical cell peripheries according to: DLVO theory (upper left), hydration repulsive forces (upper right) and corrected total energies (down). GT = total energy of interaction in DLVO approach. GHRF = energy of interaction due to hydration repulsive forces. GCT = corrected total energy of interaction. d = 2δ (δ = glycocalyx thickness). m1= principal minimum attained in a true cell contact. m2 = secondary (flocculation) minimum when the distance, r2, is attained (figure suggested by [32]).

A.I. Popescu et al. 12 1074

3.5. SPECIFIC CELLULAR INTERACTIONS

The above described cell-cell interactions do not suggest a specific interaction taking place between cells. However, there are some reasons to assume that van der Waals interactions could be specific and thus could provoke cellular segregation in a mixed culture. Indeed, if one takes into account a mixture of two types of cells, 1 and 2, with significant differences between their energies of interactions and if the below two inequalities are true:

1 1 1 2 2 2 ,VDW VDW VDWG G G↔ ↔ ↔< < (13)

it can be deduced that, the following inequality will hold, too [26, 29]:

1 1 2 2 1 22 .VDW VDW VDWG G G↔ ↔ ↔+ < (14)

The last condition is clearly describing a preferential association of the similar cells, so that the final cellular configurations will be compatible with a more favourable state from the energetic point of view. These electrodynamic forces are contributing to a cellular segregation, even in the absence of specific receptors on the cell surfaces. As concerns the repulsive electrostatic interactions, it can be deduced from Eq. (9) that these interactions are globally non-specific. However, it was suggested that these electrostatic forces could be locally attractive and slightly specific [44] due to the fact that the cell coats are also populated with minoritary positive charges.

It can be assumed that these surface charges are forming some characteristic patterns (fingerprints) of positive and negative charges. In fact, lateral diffusion of membrane components, which are carrying electrical charges, is a relatively rapid process. However, the cytoskeleton “anchorage” of these membrane components and the existence of a membrane “cortex” significantly attenuate the lateral diffusion of the membrane charged components. Therefore, at least temporarily, there is the possibility of issuing specific patterns of charges on the cell peripheries. If the presence of these specific patterns of charges is accepted, then one could admit that the associations between similar patterns are favoured in comparison with the associations between dissimilar ones, this thing pleading for a slight specific electrostatic interaction [44, 45]. Similar suggestions on the specific electrostatic attractions, due to specific ions effects [46] and preferential adhesion of the cells have been advanced [47–49]. On the other hand, the polymers dissolved, and especially, those adsorbed onto the cell surfaces can induce also some specific interactions, due to a mechanism of cross-linking, very important at short-range distances [50]. A tight contact between cells is usually mediated by specific molecules aggregated in membrane specific receptors. Therefore, the adhesion of two cells is finally performed by the binding between such specific molecules, situated on different cells, while the attachment of a cell to different partners is performed by the interactions of the membrane receptors to the partner ligands.

13 Biophysical interactions: their paramount importance for life 1075

It is interesting to note that cell-surface adhesion is determined by a quite small number of receptor species involving only 0.001 % of cell area, this meaning that rather the local surface properties than global surface ones are related to adhesive behaviour of the cells [51].

Cellular contact mediated by receptor molecules of intrinsic origin, is conferring a characteristic specificity to the cell interactions. The cell specific macromolecules in tight cell adhesion can be classified in [31, 52]: a) integrins, especially involved in cell adhesion, but some of them are involved in cell-to-cell interaction, integrins behaving alike to fibronectin, vitronectin, collagen, and laminin; b) cadherins, involved in cell-cell interactions ensuring, for instance, that cells within tissues are tight bound together; c) immunoglobulin superfamily (which are mediating the homotypic and/or heterotypic adhesion); d) selectins, single-chain transmembrane glycoproteins sharing similar properties to C-type lectins, involved cell adhesion.

3.6. DISCUSSION

The correct expression of the total energy of cellular interaction must reflect many types of energy. The interactions forces between similar and/or different cells, are mediated both by specific and non-specific interactions. The specificity of van der Waals cellular interactions is conferred by a final energetic favourable cellular configurations (conducting to cellular sorting) and, possibly, by a particular patterns of positive and negative charges on the cell peripheries (in the case of electrostatic interactions), while the strong specific interactions are mediated by intrinsic membrane receptors inserted into bilayers. The Brownian motion of cells in a culture medium or in biological fluids is, in the first moments, influenced by the long-range electrodynamic and electrostatic forces which perform a sort of preliminary selection of the interacting cells. This selection is followed by a molecular docking manoeuvre that will “force” the cells one against another favouring the operation of specific interactions mediated by membrane receptors [53–55]. Further approaching of the interacting cells is stopped by the short-range (steric and “entropic” forces) and very short-range (hydration forces).

4. CONCLUSIONS

The specific molecular interactions are driven by non specific forces: electrostatic and electrodynamic (van der Waals) forces, hydrogen bonds, salt bridges that are entering in arena during the random collisions. What is conferring to these forces a specific nature is the global 3D complementarity, completed by a

A.I. Popescu et al. 14 1076

local steric complementarity of the atoms involved in hydrogen bonding, as well as by an electrostatic complementarity of the two interacting partners. Moreover, the structural complementarity is facilitated by macromolecule chain flexibility, allowing conformational changes of the two partner molecules, according to induced fit model.

Given the occurrence of the ubiquitary specific molecular interactions, taking place in all cells, we are challenged to study the intimate mechanisms underlying such subtle interactions, in order to exploit them in view of a medical purpose.

At a superior level of living matter organisation, the cellular interactions are the result of a multitude of different molecular/macromolecular specific and non specific noncovalent interactions. Therefore the understanding of this intricate process is even a harder task.

Nowadays it is possible to adapt the technique of atomic force microscopy to directly measure the forces between individual cells [56–58]. However, much effort is still necessary to be done in order to reach a deep insight of this complex phenomenon.

A complete understanding of cell-to-cell interactions, as well as of the cell-to-substrate interactions will permit to master and influence the cells to accomplish a specific biological function and/or a desired therapeutic goal.

REFERENCES

1. C. Chilom, D. Gazdaru, N. Bucurenci, M. Straut, C. Geanta and A. Popescu, J. Optoelectronics Adv. Materials, 9, 2585–2588 (2007).

2. C. Chilom, G. Barangă, D. Găzdaru and A. Popescu, J. Optoelectronics Adv. Materials, 15, 311–316 (2013).

3. A. Popescu, S. Miron, Y. Blouquit, P. Duchambon, P. Christova and C. T. Craescu, J. Biol. Chem., 278, 40252–40261 (2003).

4. J. M Decker, http://microvet.arizona.edu/Courses/MIC419/Tutorials/ antibody.html (2007) 5. N. M. Green, Adv. Protein Chem., 29, 85–133 (1975). 6. C. G. Firanescu C. T. Craescu and A. I. Popescu, Rom. Journal of Physics, 51, 443–457 (2006). 7. A. Cooper, C. M., Johnson, J. H. Lakey, M. Nollmann, J. Biophys. Chem., 93, 215–230 (2001). 8. J. Brandts, L.-N. Lin, T. Wiseman, S. Willistonand, C. P. Yang, AM LAB, 22 (8) May, (1990). 9. L. K. Buehler, http://www.whatislife.com 10. C. Craescu, A. Bouhss J. Mispelter, E. Desis, A. Popescu, M. Chiriac and O. Barzu, 270, 7088–7096

(1995). 11. J. Dechancie, K. N. Houk, J. Am. Chem. Soc., 129, 17, 5419–29 (2007). 12. V. Raicu and A. Popescu, Integrated Molecular and Cellular Biophysics, Springer, 2008. 13. B. D. Gomperts, I. M. Kramer and P. E. R. Tatham, Signal Transduction, New York, Academic

Press (2002). 14. G. Milligan, D. Ramsay, G. Pascal and J. J. Carrillo, Life Sci., 74, 181 (2003). 15. M. C. Overton, S. L. Chinault and K. J. Blumer, Eukaryotic Cell, 4, 1963–1970 (2005). 16. G. A. Holdgate, Biotechniques, 31, 164–184 (2001). 17. A. A. Sabury, J. Chem. Thermodynamics, 35, 1975–1981 (2003). 18. J. E. Labury, Thermochimica Acta, 380, 209–215 (2001). 19. S. Leavitt, E. Freire, Curr. Op. in Chem. Biol., 11, 506–566 (2001). 20. P. C. Weber, F. R. Salemme, Curr. Op. in Structural Biol., 13, 115–121 (2003).

15 Biophysical interactions: their paramount importance for life 1077

21. V. Gheorghe, A. Popescu, Stud. Cercet. Fiz., 32, 1167–1184 (1980). 22. A. Popescu, Fundamentals of Medical Biophysics (in Romanian), Vol. II, ALL Publishing House,

Bucharest, 2001. 23. A. I. Popescu, Rom. Rep. Phys., 52, 1–2, 139–53 (2000). 24. L. D. Landau and B. V. Derjaguin, Acta Phys.-Chim. USSR, 14, 633, and J. Exp. Teor. Fiz., 11,

802 (1941). 25. E. J. W. Verwey and J. T. G. Overbeek, Theory and stability of lyophobic colloids, Elsevier,

Amsterdam, 1948. 26. V. I. Gordeliy, V. G. Cherezov, A. V. Anikin, M. V., Anikin, V. V. Ciupin and J. Texeira, Progr.

Colloid. Polym. Sci., 100, 338–344 (1996). 27. J. N. Israelachvili, Intermolecular and Surface Forces, 2nd ed., Academic Press, Jovanovich

Publishers, London, San Diego, New York, Sydney, Tokyo, Toronto, 1991. 28. V. A. Parsegian, N. Fuller and R. Rand, Proc. Natl. Acad. Sci., 76, 2750–2754 (1979). 29. R. P. Rand and V. A. Parsegian, Biochim. Biophys. Acta, 988, 351–376 (1989). 30. S. Nir, Progress in Surface Sci., 8, 1–58 (1976). 31. D. Gallez and W. T. Coakley, Heterogeneous Chemistry Rev., 3, 443–475 (1996). 32. C. J. Van Oss, Cell Biophysics, 14, 1–16 (1989). 33. W. Helfrich, Z. Naturforsch, 33a, 305–315 (1978). 34. A. K. Kenwothy, K. Hriatova, D. Needham and T. J. Mcintosh, Biophys. J. 68, 1921-1936.

(1995). 35. B. W. Ninham, Surface forces in biological systems, in: Biophysics of Water (Franks, F. ed.),

John Wiley and Sons Ltd., London, 1982, pp. 105–119. 36. A. I. Popescu and V. Gheorghe, Romanian J. Morph., Embriol. and Physiol. 26, 323–330 (1989). 37. A. C. Cowley, N. L. Fuller, R. P. Rand and V. A. Parsegian, Biochemistry, 17, 3163–3168 (1978). 38. V. A. Parsegian, R. P. Rand, N. Fuller, M. McAlister and L. J. Lis, Biorheology, 16, 293–295

(1979). 39. D. Gallez and W. T. Coakley, Prog. Biophys. Molec. Biol., 48, 155–199 (1986). 40. J. N. Israelachvili and R. M. Pashley, Biophysics of Water (Franks, F. ed.), John Wiley and Sons

Ltd., London, 1982, pp. 183–194. 41. N. V. Churaev and B. V. Derjaguin, J. Colloid. Interface. Sci., 103, 542–553 (1985). 42. T. Simonsson T. Arnebrant, L. Petersson and E. B. Hvid, Acta Odontol. Scand., 49, 311–316

(1991). 43. A. Sharma, Biophys. Chem., 47, 87–99 (1993). 44. A. I. Popescu, J. Electro- and Magnetobiol., 14, 65–74 (1995). 45. A. I. Popescu, Bioelectrochem. Bioenerg., 40, 153–157 (1996). 46. E. Ruckenstein and M. Manciu, Adv. Colloid Interface Sci., 18, 177–200 (2003) 47. H. J. Busscher, M. M. Cowan and H. C. Van der Mei, FEMS Microbiol. Rev., 8, 199–209 (1992). 48. Y. Yoshida, N. Muramatsu, K. Kataoka, H. Ohshima and T. Kondo, J. Biomater. Sci. Polym., 3,

275–285 (1992). 49. S. A. Makohliso, R. F. Valentini and P. Aebischer, J. Biomed. Mater. Res., 27, 1075–1085 (1993). 50. A. J. Baker, W. T. Coakley and D. Gallez, Eur. Biophys. J., 22, 53–62 (1993). 51. J. Vitte, A. M. Bonoliel, A. Pierres and B. Bongrand, Eur. Cell. Materials,7, 52-63 (2004). 52. B. Alberts, A. Johnson, J. Lewis, M. Raff, K. Roberts and P. Walter, Molecular Biology of the

Cell, 5-th edition, Garland Science, Taylor & Francis Group, New York, 2007. 53. G. I. Bell, Science, 200, 618–627 (1978). 54. G. I. Bell, M. Dembo and P. Bongrand, Biophys. J., 45, 1051–1064 (1984). 55. G. Cevc, Biophys. Chem., 55, 43–53 (1995). 56. I. Pera, R. Stark, M. Kappl, H.-J. Butt and F. Benefanti, Biophysical J., 87, 2446–2455 (2004). 57. J. J. Vale-Delgado, J. A. Molina-Bolivar, F. Galisteo-Gonzáles, M. J. Gálvez-Ruiz, A. Feiler and

M. W. Rutland, Langmuir, 21, 9544–9554 (2005). 58. J. Duan, J. Environ. Sci. (China), 21, 30–34 (2009).