Next Article in Journal
Chromosome 15q Structural Variants Associated with Syndromic Autism Spectrum Disorder: Clinical and Genomic Insights from Three Case Reports in a Brazilian Reference Center
Previous Article in Journal
Genome-Wide Identification and Expression of NF-YC Transcription Factors in Blueberry Under Abiotic Stress Conditions
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Opinion

Microbial Growth: Role of Water Activity and Viscoelasticity of the Cell Compartments

by
Alberto Schiraldi
Formerly at Department of Food Environmental Nutrition Sciences (DeFENS), University of Milan, 20133 Milan, Italy
Int. J. Mol. Sci. 2025, 26(17), 8508; https://doi.org/10.3390/ijms26178508
Submission received: 22 July 2025 / Revised: 26 August 2025 / Accepted: 28 August 2025 / Published: 1 September 2025
(This article belongs to the Section Molecular Biophysics)

Abstract

The complexity of the biochemistry and the variety of possible environments make the subject of the no-growth limits of bacteria a very tough challenge. This present work addresses the problem of applying to the microbial cultures the polymer science approach, which is widespread in food technology. This requires the definition of a “dynamic state diagram” that reports the expected trends of the glass transition of two virtual polymers, which mimic the crowded cytoplasmic polymers and the polymeric meshwork of the cell envelope, respectively, versus the water content. At any given temperature, the water content at the glass transition represents the lowest limit for the relevant molecular mobility. This representation leads one to recognize that the lowest temperature to observe microbial growth coincides with that of the largest freeze-concentrated liquid phase, in line with the values predicted by the Ratkowsky empirical equation. In view of potential applications in predictive microbiology, this paper suggests an alternative interpretation for the highest tolerated temperature and the temperature of the largest specific growth rate.

1. Introduction

Liquid water is ubiquitous in biological organisms and related food, pharmaceutical, and cosmetic products, where it substantially affects the overall physical properties (density, viscoelasticity, rheology, etc.), the chemical stability (oxidation, enzymatic, and non-enzymatic deterioration, etc.), and the susceptibility to microbial spoilage.
For decades, relative humidity, RH, and temperature, T, seemed to be the main physical parameters to control the detrimental effects related to moisture in food and pharmaceutical products: drying, freeze-drying, osmo-dehydration, cooling, salting, etc., became usual technological preservation practices. The relevant industrial implementation required quantitative estimation of the optimal RH and T levels for each product, or kind of product, as well as of the lowest RH and T compatible with the growth of a number of microbes [1,2,3,4,5,6], including viruses [7].
The “Labuza Map” (from the name of the pioneer in this field) reports the observed correlation between RH and the rate of many detrimental processes, including the microbial growth, at room temperature [2]. As for the effects of temperature, the so-called “Polymer Science Approach” by Slade and Levine [8] highlighted the role of the glass transition temperature, Tg, and suggested the use of a dynamical state diagram in the (T, CW) plane to define the boundaries of phases (ice, liquid, glassy/amorphous) for a water/polymer binary [9,10]. A substantial improvement of this approach came with the concept of “critical water activity”, aWC, which is the result of coupling the Tg-vs.-CW trend with the sorption isotherm [3]. In line with this view, Rahman [11] suggested enhanced stability criteria for food systems, associating the role of aW with the distance (because of either T or CW) from the glassy/rubbery state, where the molecular mobility is very poor.
On the microbiological side, Ratkowsky [12] considered the effects of T on the microbial growth. He observed that the square root of the maximum specific growth rate, μ, of most microbial cultures shows a straight-line dependence on T that allows the easy identification of Tmin as the extrapolated temperature for which μ = 0. This approach (and its enhancements; see below) is deliberately empirical and has no connection with the polymer science approach and the stability criteria mentioned above.
This present paper suggests a polymer science approach for the microbial cultures, as long as the cytoplasm and external envelope of bacterial cells show viscoelastic properties similar to those of glass-forming colloids and hydrated polymers, respectively. These properties show a correlation with the water activity of the surrounding medium. The first objective is a Tg-vs.-CW dynamic state diagram, where one can identify a viability region for microbial cells. Once coupled with the sorption isotherms of biopolymers that mimic the behavior of cytoplasm and cell envelope, this perspective becomes suitable to suggest a physical meaning for the no-growth aW and Tmin limits and justifies why both properties show values in similar ranges, namely, 0.99–1 and 260–280 K, respectively, for different kinds of microbes [2,7].

2. Viability of the Cytoplasm for a Closed Batch Microbial Culture

A viable microbial cytoplasm is not a system in thermodynamic equilibrium, since all the cytoplasmic biochemical processes require the fast displacement of small molecules, which have to interact with macromolecules (enzymes, nucleic acids, layered lipo- and gluco-proteins, polysaccharides, etc.) at specific active sites. Translational and/or segmental molecular mobility of cytoplasmic polymers within such a crowded ambient is a physical prerequisite for the viability of the cell [13,14,15,16]. If their mobility is naught, the biochemical machinery becomes shaky or inactive.
Water plays the role of “lubricant”, favoring snaking movements of the polymer chains through segmental adjustments, thanks to the disruption and reconstruction of intra- and intermolecular hydrogen bonds [17]. This is why large amounts of intracellular water “disappear”, being entrapped within the cytoplasmic polymeric entanglements that hinder easy exchanges with the surrounding bulk water, which represents a small fraction of the water content and, for polymer concentrations beyond 40% vol, experiences the electric field of the solute [14,18]. This means that the molecules of bulk water too are relented, which corresponds to a large relative viscosity, ηr, of the cytoplasm.
Because of the easy transfer of water through the cell wall, one could expect the medium and cytoplasm to have the same aW, thus implying an “osmotic” equilibrium between cells and their immediate surroundings [19]. However, this is not the case, as the microbial culture (cells + medium) is rather far from the thermodynamic equilibrium, the rates of the intracellular reactions being diffusion limited [18,20] because of the reduced mobility of both large and small molecules.
One has to look at this system from a dynamic perspective. The cell metabolism implies the transformation of SMMS (small molecular mass substrates) into new biopolymers, which will fix large amounts of water (up to ten times their own mass). This produces an extra uptake of water, mainly from the outer medium. This water flow from the medium to the cytoplasm does not correspond to a true osmotic effect but is the result of the “sequestering” action by the heavy and poorly mobile biopolymers [20,21] involved in the biochemical activity of the viable cytoplasm. The changes in the intracellular world require sophisticated experimental approaches [22,23], but the description of the extracellular aqueous medium, fortunately, is much simpler, as relatively small changes in the medium mainly concern SMMS (catabolites replace substrate, pH can vary, while some water migrates into the cells) and are easily detectable by the investigator. The difficulty to face, therefore, is to infer correct prediction of intracellular events from the reliable information about the extracellular medium.
It can be of help to approach the proposed in previous papers [24,25] that suggest a formal correlation between thermodynamic and rheological properties to account for water exchanges between a homogeneous liquid phase and a phase that shows a viscoelastic and, possibly, non-isotropic behavior. This approach treats the system with a large relative viscosity, ηr, by assigning extra terms to the chemical potential of water,
μ W = μ W * + R T l n η r α R T l n ( H ) + R T l n a W
where α ≥ 0 (for solutions of small molecules, α ≈ 0) and H accounts for the heterogeneity-related hurdles (e.g., membranes, cell walls, etc.) and has a value that depends on the direction (inbound or outbound) of the water flow.
Equation (1) predicts that the condition for no net flow of water, namely a steady state (still not a true thermodynamic equilibrium) between medium (1), where α1 ≈ 0, and cytoplasm (2) would be
a W , 1 = a W , 2   ( η r , 2 ) α 2 H 2,1 H 1,2     a W , 2   ( η r , 2 ) α 2
Now, since α2 > 0 and ηr,2 > 1, one may conclude that aW,1 > aW,2. In other words, water activity is smaller in the cytoplasm than in the outer medium, which seems consistent with the slightly lower freezing point of the cytoplasm [4,22]. It is worth noticing that this conclusion concerns just the bulk water of the cytoplasm. The water fraction sequestered by the cytoplasmic polymers does not directly contribute to the colligative properties of the cytoplasm and substantially corresponds to the so-called “unfreezable” water.
The balance between bulk and bound water fractions (which is around 50% of the overall water content of the cytoplasm, depending on the temperature and the kind of cell) seems to be the result of self-controlled feedback by the viable cells. Such feedback keeps (or quickly restores) a constant aW,2 [26], normally at the expense of the water content of the surrounding medium, whose water activity, aW,1, decreases.
It is now possible to envisage an adverse condition that compromises the viability of the cytoplasm at rest (at the temperature of the culture), namely,
a W , 1 = a W C   a W , 2   ( η r , 2 ) α 2
The cytoplasm would tend to release instead of uptake water with an increase in ηr,2, while the feedback action mentioned above would remove water molecules from some cytoplasmic macromolecules, making the polymeric entanglement less plastic, looking like a colloidal glass. Such aWC is the critical water activity (of the medium) at which the cytoplasmic polymers become too rigid to sustain the cell viability [13].
The metabolic activity of the cell ends when the flow of water ceases, namely, when the system attains the condition described by Equation (3), which therefore defines the lowest aW compatible with the cytoplasmic viability.

3. No-Growth Water Activity

Viability of cytoplasm is a necessary but not sufficient condition for cell duplication. Microbial cells can indeed survive and even produce catabolites when immobilized and/or unable to duplicate because of an external mechanical constraint [27]. A natural constraint is the stiffness of the cell wall that allows the cell to keep its shape and bear osmotic stresses up to several atmospheres. In view of the cell duplication, one has to account also for this constraint.
Vast recent literature covers the theme of the cell wall of either Gram+ or Gram− bacteria [27,28,29,30,31,32,33]. The former have an external shield of murein, which is a reticulated meshwork of peptidoglycan (PG) chains, while the latter are surrounded by an inner layer of phospholipids (PL) and an outer layer of lipopolysaccharides (LPS) with an intermediate aqueous phase (periplasm) crossed by PG thin mesh. The rigidity of either kind of shield depends on the underlying cytoplasmic biochemistry [32,33,34] that in turn depends on the molecular mobility of aqueous cytoplasmic polymers, which therefore governs also the flexibility of the cell wall.
The PG cortex behaves like a viscoelastic meshwork [27,28,29,30,31,32,33] with variable rigidity, reflected by the relevant elastic modulus, E’. It must become plastic, thanks to the selective action of hydrolases (autolysins), when the cell undergoes duplication, while it is rigid during the preceding metabolic steps and in the newborn cell, allowing water and SMMS to trespass the meshwork barrier through pores and channels. The enzymes that govern the cleavage and reconstruction of the PG meshwork come from the cytoplasmic side, as well as the aquaporins that sustain the fast aW equilibration between the hydration shell of the PG moieties, the aqueous medium, and the cytoplasmic bulk water.
In the case of Gram− bacteria, the rigidity of the outer shield mainly depends on the very tightly packed layer of the LPS bridged by Mg2+ and Ca2+ cations and changes depending on the ratio between saturated and unsaturated lipids, which is under the control of devoted cytoplasmic enzymes [34].
The focus of interest of the present work deals with the correlation between aW and viscoelastic properties of the cell polymeric components, namely, the molecular mobility of cytoplasmic macromolecules and the flexibility of the outer shield.
It is worth remembering that such correlation between water activity and rheological properties has no thermodynamic foundation [25]. However, it allows the definition of the meaningful threshold of the “critical water activity”, aWC, which, in the present case, is the value of aW of the culture medium that corresponds to the transition from large to small molecular mobility, namely, from small to large E’ of the outer shield and ηr of the cytoplasm.
In polymer science, this transition, commonly dubbed as “glass transition” or “glass-rubbery transition” actually encompasses a temperature range around the so-called glass transition temperature, Tg. In the case of aqueous polymers, this transition depends on the water mass fraction, CW = MW/M.
The so-called “dynamic state diagram” [35] reports in the plane (T, CW) the phases (liquid, ice, glassy/amorphous polymer) of an aqueous binary system. Such a diagram allows a clear view of the regions within which one may expect a large molecular mobility and then the viability of the microbial cells.
In the case of food products, the practical use of the dynamic phase diagram for systems that host a number of different polymers requires reference to a single virtual polymer with analogous rheological properties, namely, the macroscopic evidence of the Tg trend determined experimentally. To this scope, suitable experimental approaches are calorimetric (DSC) or dinamico-mechanical (DMA) investigations on samples with different CW or aW) hosted in sealed crucibles [36,37,38,39].
In the case of microbial cultures, the necessary information requires techniques that allow one to single out the glass transition of the cell interior from that of its rigid envelope. The experimental evidence must indeed concern the properties of the single cells. This makes unsuitable the results collected (with various methods) from “slurries” of cells (samples of some milligrams) [40], which therefore deal with collective properties. These are important to understand how a crowd of microbes tends to occupy the smallest volume or to form films over supporting surfaces or to investigate the survival in very low aW environments, but are useless to sketch a dynamic phase diagram.
The analysis of the incoherent elastic neutron scattering is the method proposed by Sogabe et al. [41], who related the mean square displacement (MSD) of hydrogen atoms to the molecular mobility within Cronobacter sakazakii cells. Live cell fluorescence microscopy [42] is another promising approach but still has to provide viscoelasticity data. The CLAMP (Cell Length Analysis of Mechanical Properties) technique allows determination of the elastic modulus of the cell envelope by observing the elongation of living cells entrapped within agarose gels of different stiffness [27], but does provide estimation of aWC, or Tg.
The best pieces of information come from Atomic Force Microscopy (AFM) approaches, which use a nanometric tip mounted at the end of a cantilever that mechanically solicits the cell wall and allows collection of stress relaxation data [43,44]. By applying different stresses and/or imposing different strains at various frequencies, one can determine the complex modulus, E*, and its components, E’ and E”. In simple words, gentle stresses (about 1 N) and small strains (less than 10 nm indentation of the tip) reveal the viscoelasticity of the rigid component of the cell, namely, its outer envelope, which shows the behavior of an elastic body (E’ >> E”). For larger stresses and deeper indentations of the tip, the E” component tends to prevail, as in the case of viscous fluids, just what is expected from the cytoplasm.
It turns out that the external envelopes of either Gram− or Gram+ bacteria, in spite of their chemical differences, have elastic moduli of the same order of magnitude (and are much larger than for mammalian cells) [27,45]. This supports the hypothesis that the outer membrane of Gram-negative bacteria can contribute to the overall mechanical rigidity of their envelope [34]. This means that, in the dynamic phase diagram of cultures of either Gram+ or Gram− bacteria, a single virtual polymer can represent the cell envelope, while a different virtual polymer is necessary to reflect the mechanical properties of the cytoplasm that is much more plastic than the cell envelope.
Unfortunately, the available data do not allow a quantitative definition of a Tg-vs.-CW trend. However, the widely collected evidence shows that a rigid envelope coexists with a soft cytoplasm. This means that, at the temperature of the microbial culture (say, 298 K), the former and the latter are below and above their own Tg, respectively. This allows the sketch of the expected trends (Figure 1). A single virtual polymer, P, mimics the average rheological properties of the cytoplasmic polymeric crowd, while another virtual polymer, PG, reflects the properties of the outer shield. Along a tie line at constant T, lowering CW corresponds to increasing molecular crowding (or drying) and reduced molecular mobility, while the opposite occurs when increasing CW.
For a given CW, the glass transition trend of PG, Tg,PG, runs above the Tg,P trend, i.e., Tg,PG > Tg,P. The expected slope of either Tg, for the driest conditions, is about 10 degrees Celsius per 1% increase in water mass ratio [8,9,10]. The region of the diagram between the Tg,PG and Tg,P trends corresponds to the condition of “viable” cytoplasm, as the relevant polymers acquire some mobility, being capable of short-range displacements [46,47] within a rigid capsule that does not allow cell duplication because of the mechanical constraint. The state of non-duplicating, but still viable, cells would correspond to a plastic gel entrapped in a rigid envelope, where some activity can still take place, since mobile SMMS can reach the almost immobilized endo-cellular enzymes. The cytoplasm behaves as a glass-forming liquid approaching the relevant glass transition. Its biopolymers lose their translation degrees of freedom, which damps the cell biochemical machinery [1,9,13].
This region of the dynamic state diagram encompasses the CW range smaller than the maximal freeze concentration, Cg’ (the cross point between the glass transition trend of the polymer and the decreasing trend of the cryoscopic ice separation [35]). The corresponding water, therefore, is “unfreezable”, playing just the role of conditioner of the viscoelastic properties of the cytoplasmic and wall polymers.
The CW > Cg’ region of the diagram reflects the properties of the aqueous solution of SMMS, namely, it undergoes the phase changes expected on varying the temperature, like the decreasing trend of the ice separation and the increase in boiling point.
The region of the diagram above the Tg,PG trend corresponds to the condition of more mobile cytoplasmic polymers within a flexible and modifiable outer envelope: cell duplication may take place.
The dashed area of the diagram in Figure 1 therefore represents the “Life Zone”, while the region below the Tg,P trend corresponds to physical conditions incompatible with any bioactivity.
Since microbes are dynamic systems, both cytoplasm and cell walls undergo continuous changes. This is tantamount to shifting downward or upward both Tg,P and Tg,GP, or the freezing trends (colored arrows in Figure 1).
In the case of Gram+ bacteria, the Tg,PG trend strongly depends on the degree of crosslinking between the polymeric components, regulated by specific metabolic activities [34]. This is a reaction against environmental changes [28]. An extreme metabolic reaction is sporulation [48], which implies a severe thickening of the PG shell, whose Tg can rise up to 90 °C in a P2O5 dry environment [49]. Taking into account the expected slope of the Tg trend (see above), if Tg,PG ≈ 90 °C for CW,PG ≈ 0.1 [49], one expects Tg,PG ≈ −10 °C ≡ 263 K for CW,PG ≈ 0.20.
Other metabolic adaptations concern the outer wall of Gram− microbes, where the rigidity/flexibility balance depends on the proportion between saturated and unsaturated lipid moieties of LP and LPS [34].
The lower section of Figure 1 sketches the expected desorption isotherm of a virtual biopolymer, which, for aW > 0.85 (where differences between desorption isotherms of large mass polymers become negligible for the scope of the present paper [50,51]), represents the cell polymers at the temperature of the culture. The water activity is the one of the surrounding medium (the only experimentally detectable). This plot allows the identification of the respective critical water activity, aWC,P and aWC,PG. The “no growth aW” threshold, for T = Tculture, coincides with aWC,PG, as the outer shield would become too rigid for smaller aW.
It is worth remembering that a 10 K drop of Tg for an aqueous polymer corresponds to 1% CW, which implies a very small change in its aWC. This can explain why the no-growth aW of different microbes occurs in the same narrow range of 0.99–1.
The lowest water activity of the medium to allow such residual bioactivity within the cytoplasm is the critical water activity for the virtual polymer that mimics the cytoplasmic macromolecules, aWC,P < aWC,PG. For aW < 0.8, no microbe can grow or produce signs of metabolic activity [1,2,4,5].

4. The Effect of Temperature

Figure 1 suggests another issue of some relevance. The increase in Tculture implies lower aWC and CWC, since desorption isotherms do not substantially vary for the range of aW of interest. At higher Tculture (providing that Tculture < Tmax), cellular polymers soften at lower CW, which means that the entire metabolic machinery can work in drier conditions. The effect of the temperature on the microbial growth, therefore, is a direct consequence of the change in the viscoelastic properties of the cytoplasm and cell wall.
Taking into account that viscosity and elastic modulus depend on T, showing an approximately exponential trend, exp(B/T) (with B > 0) [25], one can understand why the growth rate may appear to follow an Arrhenius trend, namely, proportional to exp(−Eatt/T), in a narrow temperature range [52]. With the aim to improve the estimations of predictive microbiology, Ratkowsky et al. [12] proposed a simple empirical alternative to the misleading Arrhenius equation, namely,
μ R a t = b T T m i n
where μ is the maximal specific growth rate, N ˙ N , N being the density of the cell population, and Tmin is the intercept on the T axis for μ = 0. As openly claimed by the authors [12], “there is no theoretical foundation for the alternative relationship to be proposed, but it does at least have the virtue of providing an excellent fit to empirical data”. Based on the above considerations and in line with some suggestions reported in the literature [53], a reasonable physical meaning for Tmin may be
Tmin = Tg,PG.
According to the growth model proposed in previous works [54,55,56,57], the maximal specific growth rate is
μ = l n ( 2 ) 3 3 8 β α
where α and β are best-fit parameters, which have a biological meaning for the virtual planktonic culture that mimics the overall growth trend of the real culture considered. In particular, β represents the theoretical maximum number of duplication steps compatible for the system (intended as medium + cells) and does not depend on the temperature (as long as the volume of the batch culture remains practically unchanged). The parameter α instead significantly depends on T, α−1/2 showing a second-order polynomial trend. This means that μ       α 1 4   T . Figure 2 shows the corresponding trend.
As for the considered psychrophiles (Figure 2), the intercept for μ = 0 corresponds to a subzero temperature, which is the case for many other microbial species [7,12,58]. In the case of the Gram+ B. subtilis [49], the values of Tmin are rather close to the expected Tg,PG (see above). There are no single-cell data about the wall of Gram-negative bacteria, but it seems reasonable to expect a substantially analogous match with the corresponding Tmin values.
With the aim to improve the predictive potential, some authors [59,60,61,62,63,64,65] extended the original Ratkowsky relationship to include two other parameters, namely, Topt and Tmax, which would correspond to the temperature of the highest μ and the maximum temperature compatible with the microbial duplication, respectively. Once again, this approach has no theoretical foundation, as it uses a single function to describe processes of different kinds. The increased number of parameters led to an enhancement of the statistical performance, although at the price of compromising the clear Ratkowsky’s original proposal. The cells of the microbial culture cannot pass from growth to no-growth condition abruptly and simultaneously. This is why one can safely predict a smooth connection between the increasing and decreasing μ trends, Tmax reflecting the thermal sustainability of the biochemical processes within the cytoplasm (and exchanges with the surrounding medium) and typically ranging between 40 and 60 °C [58,59,60,61,62,63,64,65] (where the thermally weakest cytoplasmic proteins undergo the unfolding of the native conformation). This produces the appearance of the “fake” maximum at Topt (Figure 3).
One could therefore use the μ values detected at lower temperatures to determine Tmin, according to Equation (4), and tentatively estimate Tmax and, consequently, Topt, by means of a damping function, D(T), like the one suggested by Ratkowsky et al. [58], namely,
D ( T ) = 1 e x p c T T m a x
which reflects the adverse thermal failure of the biochemical machinery. One finally gets the expression
μ = μ R a t D ( T )
that is zero for T = Tmin and T = Tmax, going through a maximum at Topt (Figure 3), although no physical, biochemical, or physiological relationship directly correlates these temperatures to one another. The search for a maximum for √μ reveals that Topt depends only on the parameter c of the damping function D(T), as the value of Topt is the root of the equation
e x p c T T m a x = c T T m i n + 1
The two functions on the left and right sides of Equation (9) are equal to 1 for T = Tmin and T = Tmax, respectively, no matter the value of the parameter c, while the related trends cross each other for T = Topt (Figure 3).
This approach implies reduced predictive performance, but avoids misleading interpretations.

5. Conclusions

The above considerations show that the medium water activity and the glass transition temperature of the polymeric components determine the overall behavior of the microbial cells, as their intertwined roles affect every process within the overcrowded intracellular environment and the rigidity of the cell wall. For this reason, aW and Tg are the preferential physical field variables to define the boundaries of the microbial viability and the cell duplication (Figure 4), substantially in line with the view of Pinto and Shimakawa [53], who recognized that the temperature-dependent bacterial growth rate is analogous to that observed in glass-forming liquids in nonliving inorganic materials.
The main conclusion of this work is that, no matter the kind and the physiology (and related biochemistry) of the microbial culture, there are common physical constraints related to the prevailing polymeric nature of the cell components, namely, aWC and Tg’, which limit cytoplasmic viability and cell duplication. Beyond these viability boundaries, only spores may exist, although their formation anyway starts within the viability region, and they return to the vegetative state once the viability conditions are restored [66,67]. This conclusion could also explain why both the no-growth aW and Tmin have values within similar ranges, namely, 0.99–1 and 260–280 K, respectively, for most microbial species with the exception of some thermophiles [6,58,59,60,61,62,63,64,65].
These limits define the boundaries of an aqueous liquid phase, the only one where molecular mobility can sustain the progress of the biochemical processes and the related irreversible production of entropy [68].

Funding

This research received no external funding.

Acknowledgments

The author has reviewed and edited the output and takes full responsibility for the content of this publication.

Conflicts of Interest

The author declares no conflicts of interest.

References

  1. Reid, D.S.; Fennema, O.R. Fennema’s Food Chemistry, 4th ed.; Damodaran, S., Parkin, K.L., Fennema, O.R., Eds.; CRC Press: Boca Raton, FL, USA, 2008; pp. 17–82. [Google Scholar]
  2. Labuza, T.P. The effect of water activity on reaction kinetics of food deterioration. Food Technol. 1980, 4, 36–59. [Google Scholar]
  3. Roos, H.Y. Phase Transitions in Foods; Academic Press Inc.: San Diego, CA, USA, 1995. [Google Scholar]
  4. Chirife, J.; Buera, M. Water activity, water glass dynamics and the control of microbiological growth in foods. Crit. Rev. Food Sci. Nutr. 1996, 36, 465–513. [Google Scholar] [CrossRef]
  5. Barbosa-Cànovas, G.V.; Fontana, A.J., Jr.; Schmidt, S.J.; Labuza, T.P. Water Activity in Foods: Fundamentals and Applications; IFT Press Series; Blackwell Publishing: Oxford, UK, 2007. [Google Scholar]
  6. Polese, P.; Del Torre, M.; Spaziani, M.; Stecchini, M.L. A simplified approach for modelling the bacterial growth/no growth boundary. Food Microbiol. 2011, 28, 384–391. [Google Scholar] [CrossRef]
  7. Roos, H.Y. Water and Pathogenic Viruses Inactivation—Food Engineering Perspectives. Food Eng. Rev. 2020, 12, 251–267. [Google Scholar] [CrossRef]
  8. Slade, L.; Levine, H. Beyond water activity: Recent advances based on an alternative approach to the assessment of food quality and safety. CRC Crit. Revs. Food Sci. Nutr. 1991, 30, 115–360. [Google Scholar] [CrossRef]
  9. Levine, H.; Slade, L. Glass Transitions in Foods. In Physical Chemistry of Foods; Schwartzberg, H.G., Hartel, R.W., Eds.; Marcel Dekker: New York, NY, USA, 1992; pp. 83–221. [Google Scholar]
  10. Roos, Y.H. Characterization of food Polymers using state diagram. J. Food Eng. 1995, 24, 339–360. [Google Scholar] [CrossRef]
  11. Rahman, M.S. Food stability determination by macro–micro region concept in the state diagram and by defining a critical temperature. J. Food Eng. 2010, 99, 402–416. [Google Scholar] [CrossRef]
  12. Ratkowsky, D.A.; Olley, J.; Mcmeekin, T.A.; Ball, A. Relationship between temperature and growth rate of bacterial cultures. J. Bacteriol. 1982, 149, 1–5. [Google Scholar] [CrossRef]
  13. Parry, B.R.; Surovtsev, I.V.; Cabeen, M.T.; O’Hern, C.S.; Dufresne, E.R.; Jacobs-Wagner, C. The bacterial cytoplasm has glass-like properties and is fluidized by metabolic activity. Cell 2014, 156, 183–194. [Google Scholar] [CrossRef]
  14. Feig, M.; Yu, I.; Wang, P.; Nawrocki, G.; Sugita, Y. Crowding in Cellular Environments at an Atomistic Level from Computer Simulations. J. Phys. Chem. B 2017, 121, 8009–8025. [Google Scholar] [CrossRef]
  15. Biology Insights Team. Intracellular Water Potential: Key Factors and Biological Impact. 2025. Available online: https://biologyinsights.com/intracellular-water-potential-key-factors-and-biological-impact/ (accessed on 29 April 2025).
  16. Lu, C.; Prada-Gracia, D.; Rao, F. Structure and dynamics of water in crowded environments slows down peptide conformational changes. J. Chem. Phys. 2014, 141, 045101. [Google Scholar] [CrossRef]
  17. Testa, A.; Dindo, M.; Rebane, A.A.; Nasouri, B.; Style, R.W.; Golestanian, R.; Dufresne, E.R.; Laurino, P. Sustained enzymatic activity and flow in crowded protein droplets. Nat. Commun. 2021, 12, 6293. [Google Scholar] [CrossRef] [PubMed]
  18. Nawrocki, G.; Sugita, Y.; Feig, M. Protein Crowding Affects Hydration Structure and Dynamics. J. Am. Chem. Soc. 2012, 134, 4842–4849. [Google Scholar]
  19. Ellis, R.J. Macromolecular crowding: Obvious but underappreciated. Trends Biochem. Sci. 2001, 26, 597–604. [Google Scholar] [CrossRef]
  20. Rivas, G.; Minton, A.P. Biochemical Reactions in Cytomimetic Media. Front. Mol. Biosci. 2020, 6, 145. [Google Scholar] [CrossRef]
  21. Verkman, A.S. Solute and macromolecule diffusion in cellular aqueous compartments. Trends Biochem. Sci. 2002, 27, 27–33. [Google Scholar] [CrossRef]
  22. Cayley, D.S.; Guttman, H.J.; Record, M.T., Jr. Biophysical characterization of changes in amounts and activity of Escherichia coli cell and compartment water and turgor pressure in response to osmotic stress. Biophys. J. 2000, 78, 1748–1764. [Google Scholar] [CrossRef]
  23. Gurian, E.; Semeraro, S.; Bellich, B.; Rampino, A.; Schneider, C.; Cesàro, A. Cell Biothermodynamics. Can calorimetry dynamically monitor cytoplasmic water activity? J. Therm. Anal. Calorim. 2017, 127, 525–534. [Google Scholar] [CrossRef]
  24. Schiraldi, A.; Fessas, D.; Signorelli, M. Water Activity in Biological Systems—A Review. Pol. J. Food Nutr. Sci. 2012, 62, 5–13. [Google Scholar] [CrossRef]
  25. Schiraldi, A. Structural Relaxation and Thermodynamics of Viscous Aqueous Systems: A Simplified Reappraisal. J. Sol. Chem. 2023, 52, 367–384. [Google Scholar] [CrossRef]
  26. Watson, J.L.; Seinkmane, E.; Styles, C.T.; Mihut, A.; Krüger, L.K.; McNally, K.E.; Planelles-Herrero, V.J.; Dudek, M.; McCall, P.M.; Barbiero, S.; et al. Macromolecular condensation buffers intracellular water potential. Nature 2023, 623, 842–885. [Google Scholar] [CrossRef]
  27. Tuson, H.H.; Auer, G.K.; Renner, L.D.; Hasebe, M.; Tropini, C.; Salick, M.; Crone, W.C.; Gopinathan, A.; Huang, K.C.; Weibel, D.B. Measuring the stiffness of bacterial cells from growth rates in hydrogels of tunable elasticity. Mol. Microbiol. 2012, 84, 874–891. [Google Scholar] [CrossRef]
  28. Borisova, M.; Gaupp, R.; Duckworth, A.; Schneider, A.; Dalügge, D.; Mühleck, M.; Deubel, D.; Unsleber, S.; Yu, W.; Muth, G.; et al. Peptidoglycan recycling in Gram-positive bacteria is crucial for survival in stationary phase. mBio 2016, 7, e00923-16. [Google Scholar] [CrossRef]
  29. Typas, A.; Banzhaf, M.; Gross, C.A.; Vollmer, W. From the regulation of peptidoglycan synthesis to bacterial growth and morphology. Nat. Rev. Microbiol. 2017, 10, 123–136. [Google Scholar] [CrossRef]
  30. Collins, T.; Margesin, R. Psychrophilic lifestyles: Mechanisms of adaptation and biotechnological tools. Appl. Microbiol. Biotechnol. 2019, 103, 2857–2871. [Google Scholar] [CrossRef]
  31. Egan, A.J.F.; Errington, J.; Vollmer, W. Regulation of peptidoglycan synthesis and remodeling. Nature 2020, 18, 446–460. [Google Scholar] [CrossRef]
  32. Wang, H.; Liu, Z.-L.; Lao, J.; Zhang, S.; Abzalimov, R.; Wang, T.; Chen, X. High Energy and Power Density Peptidoglycan Muscles through Super-Viscous Nanoconfined Water. Adv. Sci. 2022, 9, 2104697. [Google Scholar] [CrossRef]
  33. Nygaard, R.; Graham, C.L.B.; Dufrisne, M.B.; Colburn, J.D.; Pepe, J.; Hydorn, M.A.; Corradi, S.; Brown, C.M.; Ashraf, K.U.; Vickery, O.N.; et al. Structural basis of peptidoglycan synthesis by E. coli RodA-PBP2 complex. Nat. Commun. 2023, 14, 5151. [Google Scholar] [CrossRef]
  34. Sperandeo, P.; Martorana, A.M.; Polissi, A. Lipopolysaccharide biosynthesis and transport to the outer membrane of Gram-negative bacteria. In Bacterial Cell Walls and Membranes; Kuhn, A., Ed.; Springer Nature: Cham, Switzerland, 2019; Chapter 2; pp. 9–38. [Google Scholar]
  35. Corti, H.R.; Angell, C.A.; Auffret, T.; Levine, H.; Buera, M.P.; Reids, D.S.; Roos, Y.H.; Slade, L. Empirical and theoretical models of equilibrium and non-equilibrium transition temperatures of supplemented phase diagrams in aqueous systems (IUPAC Technical Report). Pure Appl. Chem. 2010, 82, 1065–1097. [Google Scholar] [CrossRef]
  36. Passot, S.; Cenard, S.; Douania, I.; Tréléa, I.C.; Fonseca, F. Critical water activity and amorphous state for optimal preservation of lyophilised lactic acid bacteria. Food Chem. 2012, 132, 1699–1705. [Google Scholar] [CrossRef]
  37. Roos, Y.H.; Karel, M. Differential Scanning Calorimetry Study of Phase Transitions Affecting the Quality of Dehydrated Materials. Biotechnol. Prog. 1990, 6, 159–163. [Google Scholar] [CrossRef]
  38. Abiad, M.G.; Campanella, O.H.; Carvajal, M.T. Assessment of Thermal Transitions by Dynamic Mechanical Analysis (DMA) Using a Novel Disposable Powder Holder. Pharmaceutics 2010, 2, 78–90. [Google Scholar] [CrossRef] [PubMed]
  39. Lee, K.; Shoda, M.; Kawai, K.; Koseki, S. Relationship between glass transition temperature, and desiccation and heat tolerance in Salmonella enterica. PLoS ONE 2020, 15, e0233638. [Google Scholar] [CrossRef]
  40. Persat, A.; Nadell, C.D.; Kim, M.K.; Ingremeau, F.; Siryaporn, A.; Drescher, K.; Wingreen, N.S.; Bassler, B.L.; Gitai, Z.; Stone, H.A. The mechanical world of bacteria. Cell 2015, 161, 988–997. [Google Scholar] [CrossRef]
  41. Sogabe, T.; Nakagawa, H.; Yamada, T.; Koseki, S.; Kawai, K. Effect of water activity on the mechanical glass transition and dynamical transition of bacteria. Biophys. J. 2022, 121, 3874–3882. [Google Scholar] [CrossRef]
  42. Hammond, L.R.; White, M.L.; Eswara, P.J. Probing Bacterial Cell Division and Cell Envelope Biogenesis with Live-Cell Fluorescence Microscopy. In The Bacterial Cell Wall. Methods in Molecular Biology; Ton-That, H., Ed.; Humana: New York, NY, USA, 2024; Volume 2727. [Google Scholar] [CrossRef]
  43. Moreno-Flores, S.; Benitez, R.; dM Vivanco, M.; Toca-Herrera, J.L. Stress relaxation and creep on living cells with the atomic force microscope: A means to calculate elastic moduli and viscosities of cell components. Nanotechnology 2010, 21, 445101. [Google Scholar] [CrossRef]
  44. Weber, A.; Tyriakowski, D.; Toca-Herrera, J.L. Power laws describe bacterial viscoelasticity. Langmuir 2022, 38, 15552–15558. [Google Scholar] [CrossRef]
  45. Pazos, M.; Peters, K. Peptidoglycan. In Bacterial Cell Walls and Membranes; Kuhn, A., Ed.; Springer Nature: Cham, Switzerland, 2019; Chapter 5; pp. 127–168. [Google Scholar]
  46. Voorhaar, L.; Hoogenboom, R. Supramolecular polymer networks: Hydrogels and bulk materials. Chem. Soc. Rev. 2016, 45, 4013–4031. [Google Scholar] [CrossRef]
  47. Bell, L.N.; Bell, H.M.; Glass, T.E. Water mobility in glassy and rubbery solids as determined by oxygen-17 nuclear magnetic resonance: Impact on chemical stability. Food Sci. Technol. 2002, 35, 108–113. [Google Scholar] [CrossRef]
  48. Yap, L.-W.; Endres, R.G. A model of cell-wall dynamics during sporulation in Bacillus subtilis. Soft Matter 2017, 13, 8089–8095. [Google Scholar] [CrossRef]
  49. Stecchini, M.L.; Del Torre, M.; Venir, E.; Morettin, A.; Furlan, P.; Maltini, E. Glassy state in Bacillus subtilis spores analyzed by differential scanning calorimetry. Int. J. Food Microbiol. 2006, 106, 286–290. [Google Scholar] [CrossRef]
  50. Pérez-Alonso, C.; Beristain, C.I.; Lobato-Calleros, C.; Rodríguez-Huezo, M.E.; Vernon-Carter, E.J. Thermodynamic analysis of the sorption isotherms of pure and blended carbohydrate polymers. J. Food Eng. 2006, 77, 753–760. [Google Scholar] [CrossRef]
  51. Berendsen, H.J.C. Specific Interactions of Water with Biopolymers. In Water in Disperse Systems; Water; Franks, F., Ed.; Springer: Boston, MA, USA, 1975; Volume 5. [Google Scholar] [CrossRef]
  52. Heitzer, A.; Kohler, H.-P.E.; Reichert, P.; Hamer, G. Utility of phenomenological models for describing temperature dependence of bacterial growth. Appl. Environ. Microbiol. 1991, 57, 2656–2665. [Google Scholar] [CrossRef] [PubMed]
  53. Pinto, L.C.; Shimakawa, K. Origin of temperature dependence of bacterial growth rate: Analogy with the viscosity of glass-forming liquids in inorganic materials. Phys. Complex Syst. 2024, 5, 67–73. [Google Scholar] [CrossRef]
  54. Schiraldi, A. The growth curve of microbial cultures: A model for a visionary reappraisal. Appl. Microbiol. 2023, 3, 288–296. [Google Scholar] [CrossRef]
  55. Schiraldi, A. The “Growth Curve”: An autocorrelation effect. Appl. Microbiol. 2024, 4, 1257–1267. [Google Scholar] [CrossRef]
  56. Schiraldi, A. The growth of microbial cultures complies with the laws of thermodynamics. Biophys. Chem. 2024, 307, 107177. [Google Scholar] [CrossRef]
  57. Schiraldi, A.; Foschino, R. A phenomenological model to infer the microbial growth: A case study for psychrotrophic pathogenic bacteria. J. Appl. Microbiol. 2022, 132, 642–653. [Google Scholar] [CrossRef]
  58. Ratkowsky, D.A.; Lowry, R.K.; Mcmeekin, T.A.; Stokes, A.N.; Chandler, R.E. Model for bacterial culture growth rate throughout the entire biokinetic temperature range. J. Bacteriol. 1983, 154, 1222–1226. [Google Scholar] [CrossRef]
  59. Rosso, L.; Lobry, R.; Bajard, S.; Flandrois, J.P. Convenient model to describe the combined effects of temperature and pH on microbial growth. Appl. Environ. Microbiol. 1995, 61, 610–616. [Google Scholar] [CrossRef]
  60. Salter, M.A.; Ross, T.; McMeekin, T.A. Applicability of a model for non-pathogenic Escherichia coli for predicting the growth of pathogenic Escherichia coli. J. Appl. Microbiol. 1998, 85, 357–364. [Google Scholar] [CrossRef]
  61. Mellefont, L.A.; McMeekin, T.A.; Ross, T. Performance evaluation of a model describing the effects of temperature, water activity, pH and lactic acid concentration on the growth of Escherichia coli. Int. J. Food Microbiol. 2003, 82, 45–58. [Google Scholar] [CrossRef]
  62. Membré, J.-M.; Leporq, B.; Vialette, M.; Mettler, E.; Perrier, L.; Thuault, D.; Zwietering, M. Temperature effect on bacterial growth rate: Quantitative microbiology approach including cardinal values and variability estimates to perform growth simulations on/in food. Int. J. Food Microbiol. 2005, 100, 179–186. [Google Scholar] [CrossRef] [PubMed]
  63. Huang, L.; Hwang, A.; Phillips, J. Effect of temperature on microbial growth rate–Mathematical analysis: The Arrhenius and Eyring–Polanyi connections. J. Food Sci. 2011, 76, 553–560. [Google Scholar] [CrossRef] [PubMed]
  64. Matejčeková, Z.; Spodniaková, S.; DujMić, E.; Liptáková, D.; Valík, L. Modelling growth of Lactobacillus plantarum as a function of temperature: Effects of media. J. Food Nutr. Res. 2019, 58, 125–134. [Google Scholar]
  65. Baranyi, J.; Buss da Silva, N.; Ellouze, M. Rethinking Tertiary Models: Relationships between Growth Parameters of Bacillus cereus Strains. Front. Microbiol. 2017, 8, 1890. [Google Scholar] [CrossRef]
  66. Gora, G.Y.; Schere, G.W.; Stone, H.A. Bacterial spores respond to humidity similarly to hydrogels. Proc. Natl. Acad. Sci. USA 2024, 121, e2320763121. [Google Scholar] [CrossRef]
  67. Atrih, A.; Foster, S.J. Analysis of the role of bacterial endospore cortex structure in resistance properties and demonstration of its conservation amongst species. J. Appl. Microbiol. 2001, 91, 364–372. [Google Scholar] [CrossRef]
  68. Schiraldi, A. Short and long-term evolution of microbial cultures. Int. J. Mol. Sci. 2025, 26, 4187. [Google Scholar] [CrossRef]
Figure 1. Upper section: Dynamic phase diagram that qualitatively reflects the properties of the aqueous fractions of a microbial cell. CW is the water mass ratio. The region with CW > Cg’ corresponds to a condition of low viscosity, just as in any aqueous solution of SMMS. The CW < Cg’ region reflects the viscoelastic behavior of the hydrated macromolecular components, namely, polymers of cytoplasm and external envelope, which are represented by two virtual polymers that mimic their glass transition trends, Tg,P and Tg,GP, respectively. Tmax is the highest temperature compatible with cell survival. Lower section: Desorption isotherm (at the temperature of the microbial culture) of the hydrated polymers P and PG, where aWC,P and aWC,PG are the corresponding critical water activity.
Figure 1. Upper section: Dynamic phase diagram that qualitatively reflects the properties of the aqueous fractions of a microbial cell. CW is the water mass ratio. The region with CW > Cg’ corresponds to a condition of low viscosity, just as in any aqueous solution of SMMS. The CW < Cg’ region reflects the viscoelastic behavior of the hydrated macromolecular components, namely, polymers of cytoplasm and external envelope, which are represented by two virtual polymers that mimic their glass transition trends, Tg,P and Tg,GP, respectively. Tmax is the highest temperature compatible with cell survival. Lower section: Desorption isotherm (at the temperature of the microbial culture) of the hydrated polymers P and PG, where aWC,P and aWC,PG are the corresponding critical water activity.
Ijms 26 08508 g001
Figure 2. The square root of the maximal specific growth rate, μ, is proportional to α−1/4 (see text), which shows a straight-line trend with an intercept at 270 K. Average values of the parameter α from the fit of growth curves of psychrophiles Aeromonas hydrophila, Listeria monocytogenes, and Yersinia enterocolitica at various temperatures [57].
Figure 2. The square root of the maximal specific growth rate, μ, is proportional to α−1/4 (see text), which shows a straight-line trend with an intercept at 270 K. Average values of the parameter α from the fit of growth curves of psychrophiles Aeromonas hydrophila, Listeria monocytogenes, and Yersinia enterocolitica at various temperatures [57].
Ijms 26 08508 g002
Figure 3. The trend of the square root of the maximal specific growth rate, μ (continuous bell-shaped blue line), goes through a maximum at T = Topt that satisfies Equation (9). The dotted blue straight line corresponds to the original Ratkowsky empirical trend. Tmin should coincide with the glass transition temperature of the outer shell, Tg,PG (see text), while Tmax is the highest temperature tolerated by the cytoplasmic biochemical machinery.
Figure 3. The trend of the square root of the maximal specific growth rate, μ (continuous bell-shaped blue line), goes through a maximum at T = Topt that satisfies Equation (9). The dotted blue straight line corresponds to the original Ratkowsky empirical trend. Tmin should coincide with the glass transition temperature of the outer shell, Tg,PG (see text), while Tmax is the highest temperature tolerated by the cytoplasmic biochemical machinery.
Ijms 26 08508 g003
Figure 4. Water activity and glass transition temperature of the polymeric cytoplasmic components and outer envelope are the field variables that define the favorable conditions for the viability and duplication of microbial cells within a narrow “viability region”. Tg’ is the lowest freeze temperature (see Figure 2). Outside this region, spores are the only surviving form of life.
Figure 4. Water activity and glass transition temperature of the polymeric cytoplasmic components and outer envelope are the field variables that define the favorable conditions for the viability and duplication of microbial cells within a narrow “viability region”. Tg’ is the lowest freeze temperature (see Figure 2). Outside this region, spores are the only surviving form of life.
Ijms 26 08508 g004
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Schiraldi, A. Microbial Growth: Role of Water Activity and Viscoelasticity of the Cell Compartments. Int. J. Mol. Sci. 2025, 26, 8508. https://doi.org/10.3390/ijms26178508

AMA Style

Schiraldi A. Microbial Growth: Role of Water Activity and Viscoelasticity of the Cell Compartments. International Journal of Molecular Sciences. 2025; 26(17):8508. https://doi.org/10.3390/ijms26178508

Chicago/Turabian Style

Schiraldi, Alberto. 2025. "Microbial Growth: Role of Water Activity and Viscoelasticity of the Cell Compartments" International Journal of Molecular Sciences 26, no. 17: 8508. https://doi.org/10.3390/ijms26178508

APA Style

Schiraldi, A. (2025). Microbial Growth: Role of Water Activity and Viscoelasticity of the Cell Compartments. International Journal of Molecular Sciences, 26(17), 8508. https://doi.org/10.3390/ijms26178508

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop