Skip to main content

ORIGINAL RESEARCH article

Front. Phys., 01 July 2022
Sec. Biophysics
This article is part of the Research Topic Active Matter in Complex Environments View all 12 articles

Biolocomotion and Premelting in Ice

  • 1Nordita, KTH Royal Institute of Technology and Stockholm University, Stockholm, Sweden
  • 2Departments of Earth and Planetary Sciences, Mathematics and Physics, Yale University, New Haven, CT, United States

Biota are found in glaciers, ice sheets and permafrost. Ice bound micro-organisms evolve in a complex mobile environment facilitated or hindered by a range of bulk and surface interactions. When a particle is embedded in a host solid near its bulk melting temperature, a melted film forms at the surface of the particle in a process known as interfacial premelting. Under a temperature gradient, the particle is driven by a thermomolecular pressure gradient toward regions of higher temperatures in a process called thermal regelation. When the host solid is ice and the particles are biota, thriving in their environment requires the development of strategies, such as producing exopolymeric substances (EPS) and antifreeze glycoproteins (AFP) that enhance the interfacial water. Therefore, thermal regelation is enhanced and modified by a process we term bio-enhanced premelting. Additionally, the motion of bioparticles is influenced by chemical gradients influenced by nutrients within the icy host body. We show how the overall trajectory of bioparticles is controlled by a competition between thermal regelation and directed biolocomotion. By re-casting this class of regelation phenomena in the stochastic framework of active Ornstein-Uhlenbeck dynamics, and using multiple scales analysis, we find that for an attractive (repulsive) nutrient source, that thermal regelation is enhanced (suppressed) by biolocomotion. This phenomena is important in astrobiology, the biosignatures of extremophiles and in terrestrial paleoclimatology.

Introduction

Ice sheets are an essential reservoir of information on past climate and they contain an important record of micro-organisms on Earth, recording ice microbes and their viruses over long periods [1, 2]. In these extreme environments, the abundance of virus is well correlated with bacterial abundance, but is 10–100 times lower than in temperate aquatic ecosystems [3]. Even in these harsh conditions, the virus infection rate is relatively high [4], leading to the expectation of low long-term survival rates. However, recent studies have shown that some viruses develop survival strategies to maintain a long-term relationship with their hosts [4, 5], possibly up to thousands of years [6]. For example, viruses such as bacteriophages can switch to a lysogenic life strategy enabling them to replicate and maintain themselves in the bacterial population without lysis over multiple generations [4]. Moreover, among these viruses some can provide immunity to their hosts against other viruses [4, 7], or manipulate their metabolism to facilitate nutrient acquisition by affecting motility genes [6]. Indeed, motile biota are found to be active in ice for substantial periods. For example, recently a 30,000 year old giant virus Pithovirus sibericum was found in permafrost along with microbes and nematodes, and viable bacteria have been found in 750,000 year old glacial ice. Basal ice often contains subglacial debris and sediment, which serve as a source of nutrients and organic matter, providing a habitat for micro-organisms adapted to subfreezing conditions [8, 9]. Additionally, the microbiomes of sediment rich basal ices are distinct from those found in glacial ice and are equivalent to those found in permafrost [8], expanding the nature of subfreezing habitats.

Ice cores provide the highest resolution records of past climate states [1015]. Of particular relevance to our study is their role as a refuge for micro-organisms, from the recent past [16, 17] to millennia [1820]. Ice microbes are taxonomically diverse and have a wide range of taxonomic relatives [9, 19, 2123]. Common algae taxa are centric and pennate diatoms, dinoflagellates and flagellates [2426], whereas common bacterial taxa are pseudomonadota, actinobacteria, firmicutes and bacteroidetes [6, 27]. Many of these microbes have different motility mechanisms [28, 29] from swimming (e.g., Chlamydomonas nivalis [30] or Methylobacterium [6, 31, 32]) to gliding (e.g., diatoms [33, 34] or Bacillus subtilis [19, 35]), which can be used to assess their locomotion. Examples of biological proxies include diatoms [36] and bacteria colonies [37, 38], reflecting a unique range of physical-biological interactions in the climate system. Therefore, understanding the relationship between the evolution of ice bound micro-organisms and proxy dating methods is a key aspect of understanding the covariation of life and climate.

Finally, such understanding is essential for the study of extraterrestrial life. In our own solar system, despite the debate regarding the existence of bulk water on Mars [39], thin water films, such as those studied here, hold the most potential for harboring life under extreme conditions. Indeed, lipids, nucleic acids, and amino acids influenced by active motility may serve as biosignatures of extra terrestial life. Combining measurements of diffusivity-characterized-motility [40, 41] with bioparticle distribution observed on Earth, provides crucial information for development of new instrumentation to detect the presence of extra terrestrial life [41, 42]. Indeed, recently micro-organisms trapped in primary fluid inclusions in halite for millions of years have been discovered [43], providing promise for both terrestrial and extraterrestrial biosignature detection.

When a particle is embedded in ice near the bulk melting temperature, the ice may melt at the particle-ice surface in a process known as interfacial premelting [44]. The thickness of the melt film depends on the temperature, impurities, material properties and geometry. A temperature gradient is accompanied by a thermomolecular pressure gradient that drives the interfacial liquid from high to low temperatures, and hence the particle migrates from low to high temperatures in a process called thermal regelation [4448]. Thermal regelation of inert particles plays a major role in the redistribution of material inside of ice, which has important environmental and composite materials implications [4448]. Moreover, surface properties are central to the fact that extremophile organisms in Earth’s cryosphere–glaciers, sea ice and permafrost–develop strategies to persist in challenging environments. Indeed, many biological organisms secrete exopolymeric substance (EPS) [49] or harness antifreeze glycoproteins (AFP) [50, 51] to maintain interfacial liquidity. For example, sea ice houses an array of algae and bacteria, some of which produce EPS to protect them at low temperature and high salinity [52, 53]. Additionally, the enhanced liquidity associated with high concentrations of EPS alters the physical properties of sea ice and thereby play a role in climate change [54, 55].

In bulk solution, active particles act as simple microscopic models for living systems and are particularly accurate at mimicking the propulsion of bacteria or algae [5660]. By converting energy to motion using biological, chemical, or physical processes, they exhibit rich collective emergent motion from ostensibly simple rules [61, 62]. Algae and bacteria operate in complex geometries and translate environmental conditions into microscopic information that guides their behavior. Examples of such information include quorum sensing (e.g., particle population density), used by bacteria to regulate biofilm formation, defense against competitors and adapt to changing environments [6365]; chemotaxis (e.g., concentration gradients of nutrients), used by algae/bacteria to direct their motion toward higher concentrations of beneficial, or lower concentrations of toxic, chemicals [6670]. It is important to emphasize that factors such as surface adhesion, salinity, the segregation of impurities of all types from the ice lattice, among other factors [67, 69, 71, 72], make our treatment of chemotaxis a simplified starting point. However, field samples and laboratory experiments have shown that cell motility is influenced by chemotaxis at low temperature [40, 73, 74]. Thus, although there are many complicated interactions that provide scope for future work, the basic role of chemotaxis in the distribution of biota in ice must start with a self-consistent framework, which is the focus of our work.

The confluence of thermal regelation, bio-enhanced premelting and intrinsic mobility underlie our study. Indeed, intrinsic mobility and chemotaxis may compete with thermal regelation, which constitutes a new area of research–ice bound active particles in premelting ice, as illustrated in Figure 1. Moreover, including micro-organism protection mechanisms that enhance interfacial liquidity, such as the secretion of EPS, constitute a unique class of regelation phenomena. Finally, treating this corpus of processes quantitatively is particularly relevant for climatology and the global carbon cycle [75, 76].

FIGURE 1
www.frontiersin.org

FIGURE 1. Schematic of few active particles embedded in ice under an external temperature gradient ∇T along the z̃-axis. The nutrient source is shown by the purple gradient. The external temperature gradient induces a drift velocity, and particles move toward regions of higher temperature, in a process known as thermal regelation (black arrow). An additional drift velocity is associated with particle motion towards higher concentration of nutrients (black dotted arrow). Thus, depending on a particle’s position, and the background temperature and nutrient gradients, these two drift effects can compete or amplify each other.

Our framework is the active Ornstein-Uhlenbeck particle (AOUP) [7783]. The active force is governed by an Ornstein-Uhlenbeck process with magnitude D̃a, which is the active diffusivity. This force can be compared to a colored noise process [79, 84]. In addition to the active diffusivity, the AOUP is characterized by a time τa, which defines the noise persistence, from which the system switches from a ballistic to a diffusive regime. The active diffusivity D̃a and characteristic time τa can be measured experimentally [8587]. The AOUP has been shown to provide accurate predictions for a range of complex phenomena [78, 80, 81, 88], and is theoretically advantageous due to its Gaussian nature [79]. These issues motivate our use of the AOUP model framework to describe the motion of active particles in ice under an external temperature gradient with a nutrient source. We analyze these particles in three dimensions using a multiple scale expansion to derive the associated Fokker-Planck equation.

The paper is organized as follows. In order to make our treatment reasonably self-contained we note that we are generalizing our previous approach [48, 89], which we recover in the appropriate limit. Thus, in §2 we introduce the active Ornstein-Uhlenbeck model for bio-premelted particles and in §3 we derive the associated Fokker-Planck equation using a multiple scale expansion. We then compare our analytic and numerical solutions after which, in §4, we draw conclusions.

Methods

Thermal regelation is understood as a consequence of the premelted film around a particle, originally treated as inert, that 1) executes diffusive motion in the ice column with diffusivity D̃(z̃)I, where I is the identity matrix, and 2) experiences a drift velocity ṽ(z̃)=U(z̃)z̃̂ parallel to the temperature gradient [48]. Therefore, regelation biases the motion of an active particle by the drift velocity U(z̃)z̃̂.

For inert particles with a sufficiently large number of moles of electrolyte impurities per unit area of surface, Ni, the premelted film thickness dNi [48, 90]. However, the production of EPS/AFP enhances liquidity at the ice surface by increasing the impurity concentration [9, 54, 72, 91], which we treat here using an enhancement factor as N = nNi, which gives

d=RgTm2NρlqmΔT,(1)

where the universal gas constant is Rg, the latent heat of fusion per mole of the solid is qm, the molar density of the liquid is ρl, the undercooling is ΔT = TmT with Tm = 273.15K the pure bulk melting temperature and T the temperature of ice.

The velocity and premelting-controlled diffusivity are given by

Uz̃=A3A231z̃3and(2)
D̃z̃=RgTmN38πνR4A23kBTmz̃3,(3)

respectively, where A2=ρlqm|T|Tm and A3=ρsqm|T|(RgTmN)36νRTm, with |∇T| the external temperature gradient. The viscosity of the fluid is ν, the particle radius is R and kB is the Boltzmann constant. Here, ρsqm ∼ 334 × 106 J m−3 [48]. The evolution of the particle position r̃=(r̃1,r̃2,r̃3)=(x̃,ỹ,z̃) and its activity are described by two overdamped Langevin equations.

ddt̃r̃t̃=βDr̃C̃r̃,t̃+2D̃aη̃+ṽz̃+2D̃z̃ξpt̃and(4)
ddt̃η̃t̃=1τaη̃t̃+1τaξat̃.(5)

The first term on the right-hand side of Eq. 4 treats the chemotaxis response, representing the effect of the nutrient source of concentration C̃ on the particle dynamics, where βD is the chemotactic strength [9295], which we treat as a constant determined by the parameters in our system. We note, however, that the transport properties of sea- and glacial-ice depend on their unique phase fraction evolution [9699], which would clearly influence the effective–porosity dependent βD. In the ideal case, wherein the nutrient source is isotropic and purely diffusive, we have

t̃C̃r̃,t̃=D̃chr̃2C̃r̃,t̃,(6)

where D̃ch is the nutrient diffusivity. The activity, or self-propulsion, is given by the term 2D̃aη̃ in Eq. 4, with D̃a the active diffusivity. The latter represents the active fluctuations of the system, such as those originating in particular processes described in Refs. [100103]. Nutrient sources, such as dissolved silica, oxygen, nitrogen and methane, play a vital role in the life of ice-bound micro-organisms, such as algae and bacteria [104109]. Here we assume that D̃ch>D̃a, consistent with [110112], and D̃ch>D̃(z̃). The function η̃=(η̃1,η̃2,η̃3) is described by an Ornstein-Uhlenbeck process, with correlations given by

ηĩt̃ηj̃t̃=δij2τae|t̃t̃|τa,(7)

where τa is the noise persistence as noted above. In the small τa limit, η̃ reduces to Gaussian white noise with correlations η̃i(t)η̃j(t)=δijδ(t̃t̃). In contrast, η̃ does not reduce to Gaussian white noise when τa is finite, and Eq. 4 does not reach equilibrium. Hence, τa controls the non-equilibrium properties of the dynamics [79, 80]. Finally, the random fluctuations in Eqs 4, 5 are given by zero mean Gaussian white noise processes ξpi(t̃)ξpj(t̃)=δijδ(t̃t̃) and ξai(t̃)ξaj(t̃)=δijδ(t̃t̃).

The Langevin Eqs 4, 5, allow us to express the probability of finding a particle at the position r̃=(r̃1,r̃2,r̃3)=(x̃,ỹ,z̃) at a given time t̃ through the Fokker-Planck equation, which describes the evolution of the probability density function P(r̃,η̃,t̃|r̃0,η̃0,t̃0), with the initial condition P(r̃,η̃,t̃=t̃0|r̃0,η̃0,t̃0)=δ(r̃r̃0)δ(η̃η̃0). To simplify the notation, we write the conditional probability as P(r̃,η̃,t̃|r̃0,η̃0,t̃0)P(r̃,η̃,t̃) and eventually arrive at the following system of coupled equations

t̃Pr̃,η̃,t̃=βDr̃Pr̃,η̃,t̃r̃C̃r̃,t̃r̃3ṽr̃3Pr̃,η̃,t̃2D̃aη̃r̃Pr̃,η̃,t̃+r̃2D̃r̃3Pr̃,η̃,t̃+1τaη̃η̃Pr̃,η̃,t̃+12τa2η̃2Pr̃,η̃,t̃and(8)
t̃C̃r̃,t̃=D̃chr̃2C̃r̃,t̃.(9)

Equations 8, 9 describe the space-time evolution of the probability of finding a particle and the concentration of nutrients respectively, akin to those of [77, 79, 94], but including the effects of thermal regelation discussed above. Both equations contain microscopic and macroscopic scales. The regime of interest is the long time behavior, computed by deriving the effective macroscopic dynamics as described next.

Results

Method of Multiple Scales

The macroscopic length characterizing the heat flux is

L=Tm|T|.(10)

The particle scale l is such that lL, and hence their ratio defines a small parameter ϵ

ϵ=lL.(11)

We use the microscopic length l and a characteristic time τ, determined a posteriori, and introduce the following dimensionless variables

η=τaη̃, r=r̃l, t=t̃τ, v=ṽu, va=ṽavac, D=D̃Dc, Dch=D̃chDn and C=C̃ch,(12)

where ṽa=2D̃aτa [113, 114], vac is the characteristic active velocity, u and Dc are the characteristic values of the regelation velocity and the premelting enhanced diffusivity respectively, and Dn and ch are the characteristic values of the diffusivity and nutrient concentration respectively. With these scalings, Eqs 8, 9, become

PltP=βDchDcrPrCPer3vPPavaηrP+r2DP+PAηηP+12PAη2P(13)
and(14)
PchtC=Dchr2C,(15)

in which we have the following dimensionless numbers,

Pe=ulDc, Pa=vaclDc, Pl=l2Dcτ, PA=l2Dcτa and Pch=l2Dnτ.(16)

We identify four characteristic time scales: tldiff=l2/Dc, tladv=l/u, tLdiff=L2/Dc and tLadv=L/u, associated with “microscopic” diffusion and advection on the particle scale, l, and “macroscopic” diffusion and advection over the thermal length scale, L. Nutrient and premelting enhanced diffusivity are taken to operate on the same time scale; tl,Ldiffntl,Ldiff. The Péclet number represents the ratio of the characteristic time for diffusion to that of advection, and those associated with regelation and activity are Pe and Pa respectively, and can be defined over both length scales,

Pe=tldifftladv and PeL=tLdifftLadv.(17)

The temperature gradient across the entire system drives thermal regelation and hence advection dominates on the macroscopic scale, so that PeL=O(1/ϵ), or equivalently, tLadv=ϵtLdiffϵtLdiffn. Whence, Pe=O(1), or equivalently, tladv=tldifftldiffn. On the macroscopic scale PeL becomes large, as ϵ tends to zero, and thus we use the macroscopic advection time τ=tLadv as our characteristic time, so that Péclet numbers based on L are O(1/ϵ) and those based on l are O(ϵ). In consequence, Eqs 13, 15 become leading to PL=O(1/ϵ) and Pl=O(ϵ), as well as PLch=O(1/ϵ) and Plch=O(ϵ). The system of Fokker-Planck equations, Eqs 1315, becomes

ϵtP=βDchDcrPrcr3vPPavaηrP+r2DP+PAηηP+12PAη2Pand(18)
ϵtC=Dchr2C.(19)

Now, we let R=r̃/L describe the macroscopic length scale, and T=t̃/tladv describe the microscopic time scale, leading to the following stretching of the microscopic scales;

r=1ϵR and T=1ϵt.(20)

Next, we use a power series ansatz for the state variables,

P=P0+ϵP1+ϵ2P2+h.o.t.and(21)
C=C0+ϵC1+ϵ2C+h.o.t.,(22)

to derive a system of equations at each order in ϵ [115], which for the concentration of nutrients, Eq. 19, are

Oϵ0:Dchr2C0=0,(23)
Oϵ1:Dchr2C1=TC02DchrRC0and(24)
Oϵ2:Dchr2C2=TC1+tC02DchrRC1DchR2C0,(25)

shown to second order. We take the approach described in [116, 117] to solve Eqs 2325. We integrate by parts over the microscale variables r and use the periodic boundary conditions to obtain the so-called weak formulation [118] of the leading order Eq. 23, the solution of which relies on the following product ansatz

C0r,R,T,t=ζrc0R,T,t.(26)

The existence and uniqueness of C0 is ensured using the Lax-Milgram theorem [118], also known as the solvability condition or the Fredholm alternative [116]. Thus, C0 is constant over C0(r,R,T,t)=C0(R,T,t). The solvability condition for the equation at O(ϵ) is

drζTc0=0,(27)

from which we find that c0 is stationary over T, leading to C0(R,T,t)=C0(R,t) and C1(R,T,t)=C1(R,t). Substituting C1 into the O(ϵ2) Eq. 25 and using the solvability condition, gives nutrient diffusion on the macroscale as

tc0=DchR2c0,(28)

showing that, as expected, the homogenization procedure is consistent with the well-known self-similar behavior of diffusion [119]. The order by order equations for the probability density function described by Eq. 18 are simplified by the observation that C0 and C1 do not depend on r, and C0 only contributes at order O(ϵ2), and hence we obtain

Oϵ0:LP0=0,(29)
Oϵ1:LP1=TP0+βDchDcrP0RC0+R3vP0+PavaηRP02rRDP0,and(30)
Oϵ2:LP2=TP1+tP0+βDchDcrP1RC1+rP0RC0+RP0RC0+R3vP1+PavaηRP12rRDP1R2DP0,(31)

where L=M+Q, with M=r3vPavaηr+r2D, and Q=PAηη+PA2η2.

Finally, as shown in Supplementary Material Section S1, upon substitution of P1 into Eq. 31 and using the solvability condition, we obtain the effective macroscopic dynamics as

t̃ρ=βDr̃ρr̃c̃z̃ṽρ+r̃2D̃a+D̃ρand(32)
t̃c̃=D̃chr̃2c̃,(33)

which are the dimensional forms of Eqs 31, 28 respectively. These capture the long time behavior wherein the active force is treated through the effective diffusivity, which is enhanced by thermal regelation, consistent with our previous work [89] and that in active matter systems generally [57, 78, 85].

Equations 32, 33 can be mapped onto the well-known Keller–Segel equations for chemotaxis [9395, 120], where ρ is the cell density and the sign of βD determines whether a cell is attracted or repelled by the nutrient. Finally, when nutrients are neglected, βD = 0, we recover our previous results [48, 89].Although Eq. 32 has an analytical solution in the large Péclet number limit, which previously allowed us to study the effect of the activity ([48, 89] or Supplementary Material Section S2), here we fix the activity and focus on the competition between thermal regelation and bio-locomotion that require solving Eqs 32, 33 numerically. We show dimensional results because of our specific interest in these processes in ice.

In the absence of nutrients, βD = 0, Figure 2 shows how the distribution of bio-particles parallel to the temperature gradient (the z̃-axis) is influenced by EPS/AFP production, which is modeled as a surface colligative effect. Namely, with Ni = 50μM m−2 and four biological enhancement factors n ∈ {1, 2, 3, 4}. The active diffusivity is D̃a=100D̃ and the particle radius is R = 9.0, ×, 10−6 m.

FIGURE 2
www.frontiersin.org

FIGURE 2. Consequences of bio-enhanced premelting in the absence of nutrients (βD = 0). The evolution of the probability density along the z̃-axis, computed from Eq. 32 with Ni = 50μM m−2 and four biological enhancement factors n ∈ {1, 2, 3, 4}, where N = nNi. The probability density is shown at x̃=ỹ=0 and at time t̃=300 years. The analytic solution (solid lines), (see Supplementary Material Section S2), is compared with the numerical solution (dots) of Eq. 32. The particle radius is R = 9.0, ×, 10−6 m.

Figure 3 shows the evolution of the nutrient concentration along the z̃-axis computed from Eq. 33, at x̃=ỹ=0. The nutrients are centered at z̃0=55m at t̃=0 and we use a nutrient diffusivity of D̃ch=1010m2s−1 [121124].

FIGURE 3
www.frontiersin.org

FIGURE 3. Evolution of the nutrient concentration (in units of M, or mol m−3) along the z̃-axis, computed from Eq. 33, at x̃=ỹ=0. At t̃=0, the nutrient concentration is centered at z̃0=55m and the nutrient diffusivity is D̃ch=1010m2s−1.

In order to study the effect of nutrients on bio-locomotion, we fix the interfacial concentration of impurities and vary the chemotaxis strength βD, where nutrients either attract (βD > 0) or repel (βD < 0) the bio-particles. Because we are interested in the situation wherein the effects of chemotaxis compete with regelation, this constrains the magnitude of βD as follows. We ask for what order of magnitude of βD are the typical chemotactic speeds approximately the same as the regelation velocity in Eq. 4. Figure 3 shows the Gaussian solution of the concentration field, with a flux that becomes arbitrarily small at long times, dominated by the algebraic contribution to r̃C̃(r̃,t̃)xt3/2exp(x2t)xt3/2. For the parameters studied here, the regelation speeds are 10−12 − 10−10m s−1 [48, 89], and hence we capture this same range in βDr̃C̃(r̃,t̃), with |βD| = 10−10 m2M−1s−1, which is realized across a large time span wherein r̃C̃(r̃,t̃) varies by several orders of magnitude. This is also reflected in the dimensionless ratio βDchDc in Eq. 13. Namely, for micron to nm scale premelted films surrounding micron sized particles Dc ranges from about 10−14 − 10−13 m2s−1, and the nutrient concentration over relevant time scales has mean values ranging over 10−3 − 10−2 M. Therefore, βDchDc ranges from 1 to 100 and hence chemotaxis is on a similar footing to regelation under these circumstances. For all cases considered here we use βD = ±10−10 m2M−1s−1 for attractive/repulsive chemotaxis.

For attractive chemotaxis (βD > 0), we show in Figure 4A the dependence of ρ(r̃,t̃) along the z̃-axis parallel to the temperature gradient and at x̃=ỹ=0, with the concentration of nutrients centered at z̃=55m. For the same conditions in the absence of chemotaxis, the net displacement from low to high temperatures due to regelation is approximately 10 m [89]. We see here the chemo-attractive modulation of ρ(r̃,t̃) during this displacement, which “pulls up” the high temperature (low z̃) tail towards the lower temperature (large z̃) but higher concentration regions centered at z̃=55m. The associated asymmetry depletes/attracts the low temperature regions at larger z̃ and concentrates the high temperature regions at smaller z̃, and is reflected in the evolution towards a sigmoidal region transecting the source at z̃=55m. As the maximum of ρ(r̃,t̃) advects through the source region it first sharpens, due to the chemo-attraction from the source “behind” it at z̃=55m, and then begins to spread out again because of the decay in the chemotactic gradient in time as seen in Figure 3 and discussed above.

FIGURE 4
www.frontiersin.org

FIGURE 4. Effect of nutrients on the particle dynamics. Evolution of the probability density function along the z̃-axis, at x̃=ỹ=0, for two values of the chemotaxis strength βD (A) βD = 10−10 m2M−1s−1 > 0 (attractive) and (B) βD = −10−10 m2M−1s−1 < 0 (repulsive). At t̃=0 the distribution is centered at z̃0=60m. The solution of Eq. 32 is computed using a finite difference method. The particle radius is R = 10−6 m, the concentration of impurities is N = 100μM m−2, the temperature gradient is |∇T| = 0.1K m−1 and the active diffusivity is D̃a=100D̃.

For repulsive chemotaxis (βD < 0), we see in Figure 4B the broad sharpening of the initial distribution in the lower temperature (large z̃) regions as it regelates/advects into the diffuse repulsive tail of nutrient field to the right of the source region centered at z̃=55m. However, because the initial high temperature (small z̃) tail of ρ(r̃,t̃) interacts with the nutrient source region at z̃=55m, chemo-repulsion quickly drives particles towards high temperature (small z̃) regions, and is clearly reflected in the creation of a local maximum. This maximum advects towards high temperature with a decaying amplitude and width due to the decay in the chemotactic gradient in time as seen in Figure 3.

In Figure 5, we show the combined effects of EPS/AFP surface enhancement of impurities in the absence of chemotaxis (βD = 0), as shown in Figure 2, and the influence of chemotaxis on particle dynamics for fixed surface impurities, as shown in Figure 4. As we vary the surface concentration of impurities we observe the same basic features as described in Figures 2, 4 and hence the same physical description applies in their interpretation. Namely, regardless of whether chemo-attraction or chemo-repulsion is operative, if the interfacial concentration of impurities N is sufficiently large then the interfacial film thicknesses are sufficiently thick that thermal regelation dominates the evolution of ρ(r̃,t̃). As the interfacial concentration of impurities N decreases chemotaxis exerts more control on the distribution, and the basic dynamics are the same as described in Figure 4. Because the magnitude of βD is fixed, and the characteristic concentration ch is 10−2 M, this N-dependence is simply assessed as discussed above, in terms of the dimensionless ratio βDchDc in Eq. 13. Namely, the numerator is fixed, but as N increases so too is the film thickness d through Eq. 1, and since Dcd3 [47], then βDchDc decreases as N−3, and the balance between chemotaxis control of the distribution gives way to regelation control. The corpus of effects studied here are reflected in this basic balance and shown in Figures 2, 4, 5.

FIGURE 5
www.frontiersin.org

FIGURE 5. The combined effects of the surface concentration of impurities and nutrients. The probability density function along the z̃-axis, at x̃=ỹ=0 and at t̃=300years, for (A) attractive (βD = 10−10 m2M−1s−1) and (B) repulsive (βD = −10−10 m2M−1s−1) chemotaxis, for different surface concentration of impurities N, computed from Eqs 32, 33 using a finite difference method. The particle radius is R = 10−6 m, the temperature gradient is |∇T| = 0.1K m−1 and the active diffusivity is D̃a=100D̃. The nutrient source is centered at z̃ = 55 m.

Conclusion

Micro-organisms in ice exhibit complex processes to persist and evolve in their harsh environments. They have developed different survival strategies, such as producing exopolymeric substances or antifreeze glycoproteins, and directing their motion toward nutrients or away from waste [34, 69, 125, 126]. We have modeled such micro-organisms using active Ornstein-Uhlenbeck particles subject to thermal regelation and biolocomotion in three dimensions. Firstly, we used a multi-scale expansion to derive the relevant coupled Fokker-Planck and diffusion Eqs 32, 33. Secondly, when nutrients are neglected, and the chemotactic strength βD = 0, we model the bio-production of surface chemicals, such as exopolymeric substances or antifreeze glycoproteins, as a surface colligative effect, and find that the associated bio-enhanced thermal regelation can dominate the distribution of particles in ice. Consistent with previous results [89], in a large Péclet number limit analytical solutions for the particle distributions are possible, and are consistent with the numerical solutions as shown in Figure 2. Thirdly, we studied the competition between thermal regelation and biolocomotion, as function of the chemotaxis strength βD, the interplay between which is shown in Figures 4, 5. The relative importance of chemo-attraction and chemo-repulsion to thermal regelation is captured by the dimensionless ratio βDchDc. When this ratio is large we find a complex modulation of regelation by chemotaxis, and when small, due to increased surface impurity concentration, leads to regelation dominated redistribution of particles. We note, however, that we have not treated the process wherein nutrients themselves have a colligative effect, which would introduce a particularly complex spatio-temporal dynamics.

Finally, we describe settings to which our analysis is applicable. It is of general interest to understand how particles in ice migrate in response to environmental forcing, as they are used as proxy to infer past climate [14, 127, 128]. Moreover, bioparticles in ice migrate in response to environmental forcing, and micro-organisms play an important role in climate change [129131]. For example, an increase in temperature activates algae/bacteria trapped in ice, producing chemicals that increase their mobility [131]. Indeed, an increase in algae/bacteria decreases the albedo of the ice [132134], thereby enhancing melting. Finally, understanding the distribution and viability of bioparticles in partially frozen media on Earth [135, 136] is essential in astrobiology [41, 42, 137].

Data Availability Statement

The original contribution presented in the study are included in the article/Supplementary Material, further inquiries can be directed to the authors.

Author Contributions

JW conceived the project. JV implemented the theory and performed simulations. JW and JV interpreted the data and wrote the paper. All authors contributed to the discussions and the final version of the manuscript.

Funding

This work was supported by the Swedish Research Council grant no. 638-2013-9243. Nordita is partially supported by Nordforsk.

Conflict of Interest

The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

Publisher’s Note

All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors and the reviewers. Any product that may be evaluated in this article, or claim that may be made by its manufacturer, is not guaranteed or endorsed by the publisher.

Acknowledgments

We thank Matthias Geilhufe, Navaneeth Marath and István Mátá Szécsényi for helpful conversations.

Supplementary Material

The Supplementary Material for this article can be found online at: https://www.frontiersin.org/articles/10.3389/fphy.2022.904836/full#supplementary-material

References

1. Karl DM, Bird DF, Björkman K, Houlihan T, Shackelford R, Tupas L. Microorganisms in the Accreted Ice of Lake Vostok, Antarctica. Science (1999) 286:2144–7. doi:10.1126/science.286.5447.2144

PubMed Abstract | CrossRef Full Text | Google Scholar

2. Christner BC, Mosley-Thompson E, Thompson LG, Reeve JN. Isolation of Bacteria and 16s rDNAs from Lake Vostok Accretion Ice. Environ Microbiol (2001) 3:570–7. doi:10.1046/j.1462-2920.2001.00226.x

PubMed Abstract | CrossRef Full Text | Google Scholar

3. Anesio AM, Mindl B, Laybourn-Parry J, Hodson AJ, Sattler B. Viral Dynamics in Cryoconite Holes on a High Arctic Glacier (Svalbard). J Geophys Res Biogeo (2007) 112:1–10. doi:10.1029/2006jg000350

CrossRef Full Text | Google Scholar

4. Bellas CM, Anesio AM, Barker G. Analysis of Virus Genomes from Glacial Environments Reveals Novel Virus Groups with Unusual Host Interactions. Front Microbiol (2015) 6:656. doi:10.3389/fmicb.2015.00656

PubMed Abstract | CrossRef Full Text | Google Scholar

5. Heilmann S, Sneppen K, Krishna S. Coexistence of Phage and Bacteria on the Boundary of Self-Organized Refuges. Proc Natl Acad Sci U.S.A (2012) 109:12828–33. doi:10.1073/pnas.1200771109

PubMed Abstract | CrossRef Full Text | Google Scholar

6. Zhong ZP, Tian F, Roux S, Gazitúa MC, Solonenko NE, Li YF, et al. Glacier Ice Archives Nearly 15,000-Year-Old Microbes and Phages. Microbiome (2021) 9:160–23. doi:10.1186/s40168-021-01106-w

PubMed Abstract | CrossRef Full Text | Google Scholar

7. Yau S, Lauro FM, DeMaere MZ, Brown MV, Thomas T, Raftery MJ, et al. Virophage Control of Antarctic Algal Host-Virus Dynamics. Proc Natl Acad Sci U.S.A (2011) 108:6163–8. doi:10.1073/pnas.1018221108

PubMed Abstract | CrossRef Full Text | Google Scholar

8. Doyle SM, Christner BC. Variation in Bacterial Composition, Diversity, and Activity Across Different Subglacial Basal Ice Types. The Cryosphere Discussions (2022) 2022:1–29. doi:10.5194/tc-2022-68

CrossRef Full Text | Google Scholar

9. Anesio AM, Lutz S, Chrismas NAM, Benning LG. The Microbiome of Glaciers and Ice Sheets. NPJ Biofilms Microbiomes (2017) 3:10–1. doi:10.1038/s41522-017-0019-0

PubMed Abstract | CrossRef Full Text | Google Scholar

10. Royer A, De Angelis M, Petit JRA. A 30000 Year Record of Physical and Optical Properties of Microparticles from an East Antarctic Ice Core and Implications for Paleoclimate Reconstruction Models. Climatic Change (1983) 5:381–412. doi:10.1007/BF00140802

CrossRef Full Text | Google Scholar

11. Legrand M, Mayewski P. Glaciochemistry of Polar Ice Cores: A Review. Rev Geophys (1997) 35:219–43. doi:10.1029/96rg03527

CrossRef Full Text | Google Scholar

12. Stauffer B, Flückiger J, Wolff E, Barnes P. The EPICA Deep Ice Cores: First Results and Perspectives. Ann Glaciol (2004) 39:93–100. doi:10.3189/172756404781814500

CrossRef Full Text | Google Scholar

13. Alley RB. Reliability of Ice-Core Science: Historical Insights. J Glaciol (2010) 56:1095–103. doi:10.3189/002214311796406130

CrossRef Full Text | Google Scholar

14. Thomas C, Ionescu D, Ariztegui D, Team DS. Impact of Paleoclimate on the Distribution of Microbial Communities in the Subsurface Sediment of the Dead Sea. Geobiology (2015) 13:546–61. doi:10.1111/gbi.12151

PubMed Abstract | CrossRef Full Text | Google Scholar

15. Tetzner D, Thomas ER, Allen CS, Wolff EW. A Refined Method to Analyze Insoluble Particulate Matter in Ice Cores, and its Application to Diatom Sampling in the Antarctic Peninsula. Front Earth Sci (2021) 9:20. doi:10.3389/feart.2021.617043

CrossRef Full Text | Google Scholar

16. Papina T, Blyakharchuk T, Eichler A, Malygina N, Mitrofanova E, Schwikowski M. Biological Proxies Recorded in a Belukha Ice Core, Russian Altai. Clim Past (2013) 9:2399–411. doi:10.5194/cp-9-2399-2013

CrossRef Full Text | Google Scholar

17. Mao G, Ji M, Xu B, Liu Y, Jiao N. Variation of High (HNA) and Low (LNA) Nucleic Acid-Content Bacteria in Tibetan Ice Cores and Their Relationship to Black Carbon. Front Microbiol (2022) 299. doi:10.3389/fmicb.2022.844432

CrossRef Full Text | Google Scholar

18. Achberger AM, Brox TI, Skidmore ML, Christner BC. Expression and Partial Characterization of an Ice Binding Protein from a Bacterium Isolated at a Depth of 3,519 Meters in the Vostok Ice Core, Antarctica. Front Microbiol (2011) 2:255. doi:10.3389/fmicb.2011.00255

PubMed Abstract | CrossRef Full Text | Google Scholar

19. Knowlton C, Veerapaneni R, D'Elia T, Rogers S. Microbial Analyses of Ancient Ice Core Sections from Greenland and Antarctica. Biology (2013) 2:206–32. doi:10.3390/biology2010206

PubMed Abstract | CrossRef Full Text | Google Scholar

20. Garcia-Lopez E, Moreno A, Bartolomé M, Leunda M, Sancho C, Cid C. Glacial Ice Age Shapes Microbiome Composition in a Receding Southern European Glacier. Front Microbiol (2021) 12:714537. doi:10.3389/fmicb.2021.714537

PubMed Abstract | CrossRef Full Text | Google Scholar

21. Wilhelm L, Singer GA, Fasching C, Battin TJ, Besemer K. Microbial Biodiversity in Glacier-Fed Streams. ISME J (2013) 7:1651–60. doi:10.1038/ismej.2013.44

PubMed Abstract | CrossRef Full Text | Google Scholar

22. Garcia-Lopez E, Moreno A, Bartolomé M, Leunda Esnaola M, Sancho C, Cid C. Glacial Ice Age Shapes Microbiome Composition in a Receding Southern European Glacier. Front Mar Sci (2021) 12. doi:10.3389/fmicb.2021.714537

CrossRef Full Text | Google Scholar

23. Stibal M, Schostag M, Cameron KA, Hansen LH, Chandler DM, Wadham JL, et al. Different Bulk and Active Bacterial Communities in Cryoconite from the Margin and interior of the Greenland Ice Sheet. Environ Microbiol Rep (2015) 7:293–300. doi:10.1111/1758-2229.12246

PubMed Abstract | CrossRef Full Text | Google Scholar

24. Hop H, Vihtakari M, Bluhm BA, Assmy P, Poulin M, Gradinger R, et al. Changes in Sea-Ice Protist Diversity with Declining Sea Ice in the Arctic Ocean from the 1980s to 2010s. Front Mar Sci (2020) 7:243. doi:10.3389/fmars.2020.00243

CrossRef Full Text | Google Scholar

25. Kauko HM, Olsen LM, Duarte P, Peeken I, Granskog MA, Johnsen G, et al. Algal Colonization of Young Arctic Sea Ice in spring. Front Mar Sci (2018) 5:199. doi:10.3389/fmars.2018.00199

CrossRef Full Text | Google Scholar

26. Spilling K, Olli K, Lehtoranta J, Kremp A, Tedesco L, Tamelander T, et al. Shifting Diatom-Dinoflagellate Dominance during Spring Bloom in the Baltic Sea and its Potential Effects on Biogeochemical Cycling. Front Mar Sci (2018) 5:327. doi:10.3389/fmars.2018.00327

CrossRef Full Text | Google Scholar

27. Itcus C, Pascu MD, Lavin P, Perşoiu A, Iancu L, Purcarea C. Bacterial and Archaeal Community Structures in Perennial Cave Ice. Sci Rep (2018) 8:15671–14. doi:10.1038/s41598-018-34106-2

PubMed Abstract | CrossRef Full Text | Google Scholar

28. Miyata M, Robinson RC, Uyeda TQP, Fukumori Y, Fukushima Si., Haruta S, et al. Tree of Motility - A Proposed History of Motility Systems in the Tree of Life. Genes Cells (2020) 25:6–21. doi:10.1111/gtc.12737

PubMed Abstract | CrossRef Full Text | Google Scholar

29. Hahnke RL, Meier-Kolthoff JP, García-López M, Mukherjee S, Huntemann M, Ivanova NN, et al. Genome-based Taxonomic Classification of Bacteroidetes. Front Microbiol (2016) 7:2003. doi:10.3389/fmicb.2016.02003

PubMed Abstract | CrossRef Full Text | Google Scholar

30. Hill NA, Häder D-P. A Biased Random Walk Model for the Trajectories of Swimming Micro-organisms. J Theor Biol (1997) 186:503–26. doi:10.1006/jtbi.1997.0421

PubMed Abstract | CrossRef Full Text | Google Scholar

31. Tsagkari E, Sloan W. The Role of the Motility of Methylobacterium in Bacterial Interactions in Drinking Water. Water (2018) 10:1386. doi:10.3390/w10101386

CrossRef Full Text | Google Scholar

32. Doerges L, Kutschera U. Assembly and Loss of the Polar Flagellum in Plant-Associated Methylobacteria. Naturwissenschaften (2014) 101:339–46. doi:10.1007/s00114-014-1162-6

PubMed Abstract | CrossRef Full Text | Google Scholar

33. Svensson F, Norberg J, Snoeijs P. Diatom Cell Size, Coloniality and Motility: Trade-Offs between Temperature, Salinity and Nutrient Supply with Climate Change. PloS One (2014) 9:e109993. doi:10.1371/journal.pone.0109993

PubMed Abstract | CrossRef Full Text | Google Scholar

34. Aumack CF, Juhl AR, Krembs C. Diatom Vertical Migration within Land-Fast Arctic Sea Ice. J Mar Syst (2014) 139:496–504. doi:10.1016/j.jmarsys.2014.08.013

CrossRef Full Text | Google Scholar

35. Christner B, Mosley-Thompson E, Thompson LG, Zagorodnov V, Sandman K, Reeve JN. Recovery and Identification of Viable Bacteria Immured in Glacial Ice. Icarus (2000) 144:479–85. doi:10.1006/icar.1999.6288

CrossRef Full Text | Google Scholar

36. Biswas RK, Choudhury AK. Diatoms: Miniscule Biological Entities with Immense Importance in Synthesis of Targeted Novel Bioparticles and Biomonitoring. J Biosci (2021) 46:1–14. doi:10.1007/s12038-021-00222-x

PubMed Abstract | CrossRef Full Text | Google Scholar

37. Dong H, Jiang H, Yu B, Liu X, Zhang C, Chan M. Impacts of Environmental Change and Human Activity on Microbial Ecosystems on the Tibetan Plateau, NW China. GSAT (2010) 20:4–10. doi:10.1130/gsatg75a.1

CrossRef Full Text | Google Scholar

38. Delgado-Baquerizo M, Bissett A, Eldridge DJ, Maestre FT, He J-Z, Wang J-T, et al. Palaeoclimate Explains a Unique Proportion of the Global Variation in Soil Bacterial Communities. Nat Ecol Evol (2017) 1:1339–47. doi:10.1038/s41559-017-0259-7

PubMed Abstract | CrossRef Full Text | Google Scholar

39. Benningfield D. The Bumpy Search for Liquid Water at the South Pole of Mars. EOS (2022) 103:1–6. doi:10.1029/2022EO220126

CrossRef Full Text | Google Scholar

40. Lindensmith CA, Rider S, Bedrossian M, Wallace JK, Serabyn E, Showalter GM, et al. A Submersible, off-axis Holographic Microscope for Detection of Microbial Motility and Morphology in Aqueous and Icy Environments. PloS one (2016) 11:e0147700. doi:10.1371/journal.pone.0147700

PubMed Abstract | CrossRef Full Text | Google Scholar

41. Nadeau J, Lindensmith C, Deming JW, Fernandez VI, Stocker R. Microbial Morphology and Motility as Biosignatures for Outer Planet Missions. Astrobiology (2016) 16:755–74. doi:10.1089/ast.2015.1376

PubMed Abstract | CrossRef Full Text | Google Scholar

42. Jones RM, Goordial JM, Orcutt BN. Low Energy Subsurface Environments as Extraterrestrial Analogs. Front Microbiol (2018) 9:1605. doi:10.3389/fmicb.2018.01605

PubMed Abstract | CrossRef Full Text | Google Scholar

43. Schreder-Gomes SI, Benison KC, Bernau JA. 830-million-year-old Microorganisms in Primary Fluid Inclusions in Halite. Geology (2022). (in press). doi:10.1130/G49957.1

CrossRef Full Text | Google Scholar

44. Dash JG, Rempel AW, Wettlaufer JS. The Physics of Premelted Ice and its Geophysical Consequences. Rev Mod Phys (2006) 78:695. doi:10.1103/RevModPhys.78.695

CrossRef Full Text | Google Scholar

45. Rempel AW, Wettlaufer JS, Worster MG. Interfacial Premelting and the Thermomolecular Force: Thermodynamic Buoyancy. Phys Rev Lett (2001) 87:088501. doi:10.1103/PhysRevLett.87.088501

PubMed Abstract | CrossRef Full Text | Google Scholar

46. Wettlaufer JS, Worster MG. Premelting Dynamics. Annu Rev Fluid Mech (2006) 38:427–52. doi:10.1146/annurev.fluid.37.061903.175758

CrossRef Full Text | Google Scholar

47. Peppin SSL, Spannuth MJ, Wettlaufer JS. Onsager Reciprocity in Premelting Solids. J Stat Phys (2009) 134:701–8. doi:10.1007/s10955-009-9699-z

CrossRef Full Text | Google Scholar

48. Marath NK, Wettlaufer JS. Impurity Effects in thermal Regelation. Soft Matter (2020) 16:5886–91. doi:10.1039/d0sm00558d

PubMed Abstract | CrossRef Full Text | Google Scholar

49. Wingender J, Neu TR, Flemming H-C. What Are Bacterial Extracellular Polymeric Substances? In: Microbial Extracellular Polymeric Substances. Berlin, Heidelberg: Springer (1999). p. 1–19. doi:10.1007/978-3-642-60147-7_1

CrossRef Full Text | Google Scholar

50. Bang JK, Lee JH, Murugan RN, Lee SG, Do H, Koh HY, et al. Antifreeze Peptides and Glycopeptides, and Their Derivatives: Potential Uses in Biotechnology. Mar Drugs (2013) 11:2013–41. doi:10.3390/md11062013

PubMed Abstract | CrossRef Full Text | Google Scholar

51. Eskandari A, Leow TC, Rahman MBA, Oslan SN. Antifreeze Proteins and Their Practical Utilization in Industry, Medicine, and Agriculture. Biomolecules (2020) 10:1649. doi:10.3390/biom10121649

CrossRef Full Text | Google Scholar

52. Ewert M, Deming JW. Bacterial Responses to Fluctuations and Extremes in Temperature and Brine Salinity at the Surface of Arctic winter Sea Ice. FEMS Microbiol Ecol (2014) 89:476–89. doi:10.1111/1574-6941.12363

PubMed Abstract | CrossRef Full Text | Google Scholar

53. Ewert M, Deming JW. Sea Ice Microorganisms: Environmental Constraints and Extracellular Responses. Biology (Basel) (2013) 2:603–28. doi:10.3390/biology2020603

PubMed Abstract | CrossRef Full Text | Google Scholar

54. Krembs C, Eicken H, Deming JW. Exopolymer Alteration of Physical Properties of Sea Ice and Implications for Ice Habitability and Biogeochemistry in a Warmer Arctic. Proc Natl Acad Sci U.S.A (2011) 108:3653–8. doi:10.1073/pnas.1100701108

PubMed Abstract | CrossRef Full Text | Google Scholar

55. Decho AW, Gutierrez T. Microbial Extracellular Polymeric Substances (EPSs) in Ocean Systems. Front Microbiol (2017) 8:922. doi:10.3389/fmicb.2017.00922

PubMed Abstract | CrossRef Full Text | Google Scholar

56. Cates ME. Diffusive Transport without Detailed Balance in Motile Bacteria: Does Microbiology Need Statistical Physics? Rep Prog Phys (2012) 75:042601. doi:10.1088/0034-4885/75/4/042601

PubMed Abstract | CrossRef Full Text | Google Scholar

57. Bechinger C, Di Leonardo R, Löwen H, Reichhardt C, Volpe G, Volpe G. Active Particles in Complex and Crowded Environments. Rev Mod Phys (2016) 88:045006. doi:10.1103/revmodphys.88.045006

CrossRef Full Text | Google Scholar

58. Ghosh A, Fischer P. Controlled Propulsion of Artificial Magnetic Nanostructured Propellers. Nano Lett (2009) 9:2243–5. doi:10.1021/nl900186w

PubMed Abstract | CrossRef Full Text | Google Scholar

59. Kim S, Qiu F, Kim S, Ghanbari A, Moon C, Zhang L, et al. Fabrication and Characterization of Magnetic Microrobots for Three‐Dimensional Cell Culture and Targeted Transportation. Adv Mater (2013) 25:5863–8. doi:10.1002/adma.201301484

PubMed Abstract | CrossRef Full Text | Google Scholar

60. Jin C, Vachier J, Bandyopadhyay S, Macharashvili T, Maass CC. Fine Balance of Chemotactic and Hydrodynamic Torques: When Microswimmers Orbit a Pillar Just once. Phys Rev E (2019) 100:040601. doi:10.1103/PhysRevE.100.040601

PubMed Abstract | CrossRef Full Text | Google Scholar

61. Elgeti J, Winkler RG, Gompper G. Physics of Microswimmers-Single Particle Motion and Collective Behavior: A Review. Rep Prog Phys (2015) 78:056601. doi:10.1088/0034-4885/78/5/056601

PubMed Abstract | CrossRef Full Text | Google Scholar

62. Romanczuk P, Bär M, Ebeling W, Lindner B, Schimansky-Geier L. Active Brownian Particles. Eur Phys J Spec Top (2012) 202:1–162. doi:10.1140/epjst/e2012-01529-y

CrossRef Full Text | Google Scholar

63. Li Y-H, Tian X. Quorum sensing and Bacterial Social Interactions in Biofilms. Sensors (2012) 12:2519–38. doi:10.3390/s120302519

PubMed Abstract | CrossRef Full Text | Google Scholar

64. Yan S, Wu G. Can Biofilm Be Reversed through Quorum Sensing in Pseudomonas aeruginosa? Front Microbiol (2019) 10:1582. doi:10.3389/fmicb.2019.01582

PubMed Abstract | CrossRef Full Text | Google Scholar

65. Lee CK, Vachier J, de Anda J, Zhao K, Baker AE, Bennett RR, et al. Social Cooperativity of Bacteria during Reversible Surface Attachment in Young Biofilms: A Quantitative Comparison of Pseudomonas aeruginosa PA14 and PAO1. MBio (2020) 11:e02644–19. doi:10.1128/mBio.02644-19

PubMed Abstract | CrossRef Full Text | Google Scholar

66. Wadhams GH, Armitage JP. Making Sense of it All: Bacterial Chemotaxis. Nat Rev Mol Cel Biol (2004) 5:1024–37. doi:10.1038/nrm1524

CrossRef Full Text | Google Scholar

67. Showalter GM, Deming JW. Low-temperature Chemotaxis, Halotaxis and Chemohalotaxis by the Psychrophilic marine Bacterium Colwellia Psychrerythraea 34h. Environ Microbiol Rep (2018) 10:92–101. doi:10.1111/1758-2229.12610

PubMed Abstract | CrossRef Full Text | Google Scholar

68. Cremer J, Honda T, Tang Y, Wong-Ng J, Vergassola M, Hwa T. Chemotaxis as a Navigation Strategy to Boost Range Expansion. Nature (2019) 575:658–63. doi:10.1038/s41586-019-1733-y

PubMed Abstract | CrossRef Full Text | Google Scholar

69. Bar Dolev M, Bernheim R, Guo S, Davies PL, Braslavsky I. Putting Life on Ice: Bacteria that Bind to Frozen Water. J R Soc Interf (2016) 13:20160210. doi:10.1098/rsif.2016.0210

CrossRef Full Text | Google Scholar

70. Mattingly H, Kamino K, Machta B, Emonet T. Escherichia coli Chemotaxis is Information Limited. Nat Phys (2021) 17:1–6. doi:10.1038/s41567-021-01380-3

CrossRef Full Text | Google Scholar

71. Bar-Dolev M, Celik Y, Wettlaufer JS, Davies PL, Braslavsky I. New Insights into Ice Growth and Melting Modifications by Antifreeze Proteins. J R Soc Interf (2012) 9:3249–59. doi:10.1098/rsif.2012.0388

CrossRef Full Text | Google Scholar

72. Hansen-Goos H, Thomson ES, Wettlaufer JS. On the Edge of Habitability and the Extremes of Liquidity. Planet Space Sci (2014) 98:169–81. doi:10.1016/j.pss.2013.04.010

CrossRef Full Text | Google Scholar

73. Junge K, Eicken H, Deming JW. Motility of Colwellia Psychrerythraea Strain 34H at Subzero Temperatures. Appl Environ Microbiol (2003) 69:4282–4. doi:10.1128/aem.69.7.4282-4284.2003

PubMed Abstract | CrossRef Full Text | Google Scholar

74. Mudge MC, Nunn BL, Firth E, Ewert M, Hales K, Fondrie WE, et al. Subzero, saline Incubations of Colwellia Psychrerythraea Reveal Strategies and Biomarkers for Sustained Life in Extreme Icy Environments. Environ Microbiol (2021) 23:3840–66. doi:10.1111/1462-2920.15485

PubMed Abstract | CrossRef Full Text | Google Scholar

75. Wadham JL, Hawkings JR, Tarasov L, Gregoire LJ, Spencer RGM, Gutjahr M, et al. Ice Sheets Matter for the Global Carbon Cycle. Nat Commun (2019) 10:3567–17. doi:10.1038/s41467-019-11394-4

PubMed Abstract | CrossRef Full Text | Google Scholar

76. Holland AT, Williamson CJ, Sgouridis F, Tedstone AJ, McCutcheon J, Cook JM, et al. Dissolved Organic Nutrients Dominate Melting Surface Ice of the Dark Zone (Greenland Ice Sheet). Biogeosciences (2019) 16:3283–96. doi:10.5194/bg-16-3283-2019

CrossRef Full Text | Google Scholar

77. Fodor É, Nardini C, Cates ME, Tailleur J, Visco P, van Wijland F. How Far from Equilibrium is Active Matter? Phys Rev Lett (2016) 117:038103. doi:10.1103/PhysRevLett.117.038103

PubMed Abstract | CrossRef Full Text | Google Scholar

78. Caprini L, Marini Bettolo Marconi U. Active Particles under Confinement and Effective Force Generation Among Surfaces. Soft Matter (2018) 14:9044–54. doi:10.1039/c8sm01840e

PubMed Abstract | CrossRef Full Text | Google Scholar

79. Martin D, O'Byrne J, Cates ME, Fodor É, Nardini C, Tailleur J, et al. Statistical Mechanics of Active Ornstein-Uhlenbeck Particles. Phys Rev E (2021) 103:032607. doi:10.1103/PhysRevE.103.032607

PubMed Abstract | CrossRef Full Text | Google Scholar

80. Dabelow L, Eichhorn R. Irreversibility in Active Matter: General Framework for Active Ornstein-Uhlenbeck Particles. Front Phys (2021) 8:516. doi:10.3389/fphy.2020.582992

CrossRef Full Text | Google Scholar

81. Caprini L, Marini Bettolo Marconi U, Puglisi A. Activity Induced Delocalization and Freezing in Self-Propelled Systems. Sci Rep (2019) 9:1386–9. doi:10.1038/s41598-018-36824-z

PubMed Abstract | CrossRef Full Text | Google Scholar

82. Bonilla LL. Active Ornstein-Uhlenbeck Particles. Phys Rev E (2019) 100:022601. doi:10.1103/PhysRevE.100.022601

PubMed Abstract | CrossRef Full Text | Google Scholar

83. Dabelow L, Bo S, Eichhorn R. Irreversibility in Active Matter Systems: Fluctuation Theorem and Mutual Information. Phys Rev X (2019) 9:021009. doi:10.1103/physrevx.9.021009

CrossRef Full Text | Google Scholar

84. Sevilla FJ, Rodríguez RF, Gomez-Solano JR. Generalized Ornstein-Uhlenbeck Model for Active Motion. Phys Rev E (2019) 100:032123. doi:10.1103/PhysRevE.100.032123

PubMed Abstract | CrossRef Full Text | Google Scholar

85. Maggi C, Paoluzzi M, Pellicciotta N, Lepore A, Angelani L, Di Leonardo R. Generalized Energy Equipartition in Harmonic Oscillators Driven by Active Baths. Phys Rev Lett (2014) 113:238303. doi:10.1103/physrevlett.113.238303

PubMed Abstract | CrossRef Full Text | Google Scholar

86. Maggi C, Paoluzzi M, Angelani L, Di Leonardo R. Memory-less Response and Violation of the Fluctuation-Dissipation Theorem in Colloids Suspended in an Active bath. Sci Rep (2017) 7:17588–7. doi:10.1038/s41598-017-17900-2

PubMed Abstract | CrossRef Full Text | Google Scholar

87. Donado F, Moctezuma RE, López-Flores L, Medina-Noyola M, Arauz-Lara JL. Brownian Motion in Non-equilibrium Systems and the Ornstein-Uhlenbeck Stochastic Process. Sci Rep (2017) 7:12614–7. doi:10.1038/s41598-017-12737-1

PubMed Abstract | CrossRef Full Text | Google Scholar

88. Marini Bettolo Marconi U, Maggi C, Melchionna S. Pressure and Surface Tension of an Active Simple Liquid: A Comparison between Kinetic, Mechanical and Free-Energy Based Approaches. Soft Matter (2016) 12:5727–38. doi:10.1039/c6sm00667a

PubMed Abstract | CrossRef Full Text | Google Scholar

89. Vachier J, Wettlaufer JS. Premelting Controlled Active Matter in Ice. Phys Rev E (2022) 105:024601. doi:10.1103/PhysRevE.105.024601

PubMed Abstract | CrossRef Full Text | Google Scholar

90. Wettlaufer JS. Impurity Effects in the Premelting of Ice. Phys Rev Lett (1999) 82:2516–9. doi:10.1103/physrevlett.82.2516

CrossRef Full Text | Google Scholar

91. Ewert M, Deming J. Sea Ice Microorganisms: Environmental Constraints and Extracellular Responses. Biology (2013) 2:603–28. doi:10.3390/biology2020603

PubMed Abstract | CrossRef Full Text | Google Scholar

92. Keller EF, Segel LA. Model for Chemotaxis. J Theor Biol (1971) 30:225–34. doi:10.1016/0022-5193(71)90050-6

PubMed Abstract | CrossRef Full Text | Google Scholar

93. Saha S, Golestanian R, Ramaswamy S. Clusters, Asters, and Collective Oscillations in Chemotactic Colloids. Phys Rev E Stat Nonlin Soft Matter Phys (2014) 89:062316. doi:10.1103/PhysRevE.89.062316

PubMed Abstract | CrossRef Full Text | Google Scholar

94. Liebchen B, Löwen H. Synthetic Chemotaxis and Collective Behavior in Active Matter. Acc Chem Res (2018) 51:2982–90. doi:10.1021/acs.accounts.8b00215

PubMed Abstract | CrossRef Full Text | Google Scholar

95. Pohl O, Stark H. Dynamic Clustering and Chemotactic Collapse of Self-Phoretic Active Particles. Phys Rev Lett (2014) 112:238303. doi:10.1103/physrevlett.112.238303

PubMed Abstract | CrossRef Full Text | Google Scholar

96. Wettlaufer JS, Worster MG, Huppert HE. Natural Convection during Solidification of an alloy from above with Application to the Evolution of Sea Ice. J Fluid Mech (1997) 344:291–316. doi:10.1017/s0022112097006022

CrossRef Full Text | Google Scholar

97. Rempel AW, Waddington ED, Wettlaufer JS, Worster MG. Possible Displacement of the Climate Signal in Ancient Ice by Premelting and Anomalous Diffusion. Nature (2001) 411:568–71. doi:10.1038/35079043

PubMed Abstract | CrossRef Full Text | Google Scholar

98. Rempel AW, Wettlaufer JS, Waddington ED. Anomalous Diffusion of Multiple Impurity Species: Predicted Implications for the Ice Core Climate Records. J Geophys Res Solid Earth (2002) 107:ECV 3–1–ECV 3–12. doi:10.1029/2002jb001857

CrossRef Full Text | Google Scholar

99. Wells AJ, Wettlaufer JS, Orszag SA. Maximal Potential Energy Transport: A Variational Principle for Solidification Problems. Phys Rev Lett (2010) 105:254502. doi:10.1103/physrevlett.105.254502

PubMed Abstract | CrossRef Full Text | Google Scholar

100. Joanny J-F, Jülicher F, Prost J. Motion of an Adhesive Gel in a Swelling Gradient: A Mechanism for Cell Locomotion. Phys Rev Lett (2003) 90:168102. doi:10.1103/physrevlett.90.168102

PubMed Abstract | CrossRef Full Text | Google Scholar

101. Peruani F, Morelli LG. Self-propelled Particles with Fluctuating Speed and Direction of Motion in Two Dimensions. Phys Rev Lett (2007) 99:010602. doi:10.1103/PhysRevLett.99.010602

PubMed Abstract | CrossRef Full Text | Google Scholar

102. Romanczuk P, Schimansky-Geier L. Brownian Motion with Active Fluctuations. Phys Rev Lett (2011) 106:230601. doi:10.1103/physrevlett.106.230601

PubMed Abstract | CrossRef Full Text | Google Scholar

103. Vandebroek H, Vanderzande C. The Effect of Active Fluctuations on the Dynamics of Particles, Motors and DNA-Hairpins. Soft Matter (2017) 13:2181–91. doi:10.1039/c6sm02568d

PubMed Abstract | CrossRef Full Text | Google Scholar

104. Price PB. A Habitat for Psychrophiles in Deep Antarctic Ice. Proc Natl Acad Sci U.S.A (2000) 97:1247–51. doi:10.1073/pnas.97.3.1247

PubMed Abstract | CrossRef Full Text | Google Scholar

105. Campen RK, Sowers T, Alley RB. Evidence of Microbial Consortia Metabolizing within a Low-Latitude Mountain Glacier. Geol (2003) 31:231–4. doi:10.1130/0091-7613(2003)031<0231:eomcmw>2.0.co;2

CrossRef Full Text | Google Scholar

106. Tung HC, Price PB, Bramall NE, Vrdoljak G. Microorganisms Metabolizing on clay Grains in 3-Km-Deep Greenland Basal Ice. Astrobiology (2006) 6:69–86. doi:10.1089/ast.2006.6.69

PubMed Abstract | CrossRef Full Text | Google Scholar

107. Mader HM, Pettitt ME, Wadham JL, Wolff EW, Parkes RJ. Subsurface Ice as a Microbial Habitat. Geol (2006) 34:169–72. doi:10.1130/g22096.1

CrossRef Full Text | Google Scholar

108. Rohde RA, Price PB. Diffusion-controlled Metabolism for Long-Term Survival of Single Isolated Microorganisms Trapped within Ice Crystals. Proc Natl Acad Sci U.S.A (2007) 104:16592–7. doi:10.1073/pnas.0708183104

PubMed Abstract | CrossRef Full Text | Google Scholar

109. Vancoppenolle M, Goosse H, De Montety A, Fichefet T, Tremblay B, Tison JL. Modeling Brine and Nutrient Dynamics in Antarctic Sea Ice: The Case of Dissolved Silica. J Geophys Res Oceans (2010) 115:1–18. doi:10.1029/2009jc005369

CrossRef Full Text | Google Scholar

110. Wu M, Roberts JW, Kim S, Koch DL, DeLisa MP. Collective Bacterial Dynamics Revealed Using a Three-Dimensional Population-Scale Defocused Particle Tracking Technique. Appl Environ Microbiol (2006) 72:4987–94. doi:10.1128/aem.00158-06

PubMed Abstract | CrossRef Full Text | Google Scholar

111. Ben Amar M. Collective Chemotaxis and Segregation of Active Bacterial Colonies. Sci Rep (2016) 6:21269–9. doi:10.1038/srep21269

PubMed Abstract | CrossRef Full Text | Google Scholar

112. Wåhlin J, Klein-Paste A. The Effect of Mass Diffusion on the Rate of Chemical Ice Melting Using Aqueous Solutions. Cold Regions Sci Techn (2017) 139:11–21. doi:10.1016/j.coldregions.2017.04.001

CrossRef Full Text | Google Scholar

113. Dabelow L, Bo S, Eichhorn R. How Irreversible Are Steady-State Trajectories of a Trapped Active Particle? J Stat Mech Theor E. (2021) 2021:033216. doi:10.1088/1742-5468/abe6fd

CrossRef Full Text | Google Scholar

114. Caprini L, Marconi UMB, Wittmann R, Löwen H. Dynamics of Active Particles with Space-dependent Swim Velocity. Soft Matter (2022) 18:1412–22. doi:10.1039/d1sm01648b

PubMed Abstract | CrossRef Full Text | Google Scholar

115. Bender CM, Orszag SA. Advanced Mathematical Methods for Scientists and Engineers: Asymptotic Methods and Perturbation Theory. New York, NY: Springer (2013).

Google Scholar

116. Pavliotis G, Stuart A. Multiscale Methods: Averaging and Homogenization. New York, NY: Springer (2008).

Google Scholar

117. Aurell E, Bo S, Dias M, Eichhorn R, Marino R. Diffusion of a Brownian Ellipsoid in a Force Field. EPL (2016) 114:30005. doi:10.1209/0295-5075/114/30005

CrossRef Full Text | Google Scholar

118. Chipot M. Weak Formulation of Elliptic Problems. In: Elliptic Equations: An Introductory Course. Basel: Birkhäuser, Basel (2009). p. 35–42. doi:10.1007/978-3-7643-9982-5_3

CrossRef Full Text | Google Scholar

119. Barenblatt GI. Scaling, Self-Similarity, and Intermediate Asymptotics. Cambridge, UK: Cambridge University Press (1996).

Google Scholar

120. Liebchen B, Löwen H. Modeling Chemotaxis of Microswimmers: From Individual to Collective Behavior. In: Chemical Kinetics: Beyond the Textbook. New Jersey: World Scientific (2020). p. 493–516.

Google Scholar

121. Theurkauff I, Cottin-Bizonne C, Palacci J, Ybert C, Bocquet L. Dynamic Clustering in Active Colloidal Suspensions with Chemical Signaling. Phys Rev Lett (2012) 108:268303. doi:10.1103/physrevlett.108.268303

PubMed Abstract | CrossRef Full Text | Google Scholar

122. Jin C, Krüger C, Maass CC. Chemotaxis and Autochemotaxis of Self-Propelling Droplet Swimmers. Proc Natl Acad Sci U.S.A (2017) 114:5089–94. doi:10.1073/pnas.1619783114

PubMed Abstract | CrossRef Full Text | Google Scholar

123. Salek MM, Carrara F, Fernandez V, Guasto JS, Stocker R. Bacterial Chemotaxis in a Microfluidic T-Maze Reveals strong Phenotypic Heterogeneity in Chemotactic Sensitivity. Nat Commun (2019) 10:1877–11. doi:10.1038/s41467-019-09521-2

PubMed Abstract | CrossRef Full Text | Google Scholar

124. Hokmabad BV, Agudo-Canalejo J, Saha S, Golestanian R, Maass CC. Chemotactic Self-Caging in Active Emulsions (2022) 119 (24):e2122269119. doi:10.1073/pnas.2122269119

PubMed Abstract | CrossRef Full Text | Google Scholar

125. Price PB. Microbial Life in Glacial Ice and Implications for a Cold Origin of Life. FEMS Microbiol Ecol (2007) 59:217–31. doi:10.1111/j.1574-6941.2006.00234.x

PubMed Abstract | CrossRef Full Text | Google Scholar

126. Stocker R, Seymour JR. Ecology and Physics of Bacterial Chemotaxis in the Ocean. Microbiol Mol Biol Rev (2012) 76:792–812. doi:10.1128/mmbr.00029-12

PubMed Abstract | CrossRef Full Text | Google Scholar

127. Miteva V. Bacteria in Snow and Glacier Ice. In: Psychrophiles: From Biodiversity to Biotechnology. Berlin, Heidelberg: Springer (2008). p. 31–50. doi:10.1007/978-3-540-74335-4_3

CrossRef Full Text | Google Scholar

128. Han D, Nam SI, Kim JH, Stein R, Niessen F, Joe YJ, et al. Inference on Paleoclimate Change Using Microbial Habitat Preference in Arctic Holocene Sediments. Sci Rep (2017) 7:9652–11. doi:10.1038/s41598-017-08757-6

PubMed Abstract | CrossRef Full Text | Google Scholar

129. Mitchell JG, Kogure K. Bacterial Motility: Links to the Environment and a Driving Force for Microbial Physics. FEMS Microbiol Ecol (2006) 55:3–16. doi:10.1111/j.1574-6941.2005.00003.x

PubMed Abstract | CrossRef Full Text | Google Scholar

130. Dutta H, Dutta A. The Microbial Aspect of Climate Change. Energ Ecol Environ (2016) 1:209–32. doi:10.1007/s40974-016-0034-7

CrossRef Full Text | Google Scholar

131. Cavicchioli R, Ripple WJ, Timmis KN, Azam F, Bakken LR, Baylis M, et al. Scientists' Warning to Humanity: Microorganisms and Climate Change. Nat Rev Microbiol (2019) 17:569–86. doi:10.1038/s41579-019-0222-5

PubMed Abstract | CrossRef Full Text | Google Scholar

132. Ryan JC, Hubbard A, Stibal M, Irvine-Fynn TD, Cook J, Smith LC, et al. Dark Zone of the Greenland Ice Sheet Controlled by Distributed Biologically-Active Impurities. Nat Commun (2018) 9:1065–10. doi:10.1038/s41467-018-03353-2

PubMed Abstract | CrossRef Full Text | Google Scholar

133. Perini L, Gostinčar C, Anesio AM, Williamson C, Tranter M, Gunde-Cimerman N. Darkening of the Greenland Ice Sheet: Fungal Abundance and Diversity Are Associated with Algal Bloom. Front Microbiol (2019) 10:557. doi:10.3389/fmicb.2019.00557

PubMed Abstract | CrossRef Full Text | Google Scholar

134. Williamson CJ, Cook J, Tedstone A, Yallop M, McCutcheon J, Poniecka E, et al. Algal Photophysiology Drives Darkening and Melt of the Greenland Ice Sheet. Proc Natl Acad Sci U.S.A (2020) 117:5694–705. doi:10.1073/pnas.1918412117

PubMed Abstract | CrossRef Full Text | Google Scholar

135. Van Leeuwe MA, Tedesco L, Arrigo KR, Assmy P, Campbell K, Meiners KM, et al. Microalgal Community Structure and Primary Production in Arctic and Antarctic Sea Ice: A Synthesis. Element Sci Anthropocene (2018) 6:1–25. doi:10.1525/elementa.267

CrossRef Full Text | Google Scholar

136. Cimoli E, Lucieer V, Meiners KM, Chennu A, Castrisios K, Ryan KG, et al. Mapping the In Situ Microspatial Distribution of Ice Algal Biomass through Hyperspectral Imaging of Sea-Ice Cores. Sci Rep (2020) 10:21848–17. doi:10.1038/s41598-020-79084-6

PubMed Abstract | CrossRef Full Text | Google Scholar

137. Wettlaufer JS. Sea Ice and Astrobiology. In: D Thomas, and G Dieckmann, editors. Sea Ice. 2nd ed. Oxford: Wiley-Blackwell (2010). p. 579–94. chap. 15.

Google Scholar

138. Risken H. Fokker-Planck Equation. In: The Fokker-Planck Equation. Berlin, Heidelberg: Springer (1996). p. 63–95. doi:10.1007/978-3-642-61544-3_4

CrossRef Full Text | Google Scholar

139. Celani A, Vergassola M. Bacterial Strategies for Chemotaxis Response. Proc Natl Acad Sci U.S.A (2010) 107:1391–6. doi:10.1073/pnas.0909673107

PubMed Abstract | CrossRef Full Text | Google Scholar

140. Kheifets S. Application of the Green’s Function Method to Some Nonlinear Problems of an Electron Storage Ring: Part 1, the Green’s Function for the Fokker-Planck Equation. Technical Report. Menlo Park, CA (USA): Stanford Linear Accelerator Center (1982).

Google Scholar

Keywords: bioparticles, premelting, biolocomotion, active matter, Ornstein-Uhlenbeck process, extremophiles

Citation: Vachier  J and Wettlaufer  JS (2022) Biolocomotion and Premelting in Ice. Front. Phys. 10:904836. doi: 10.3389/fphy.2022.904836

Received: 25 March 2022; Accepted: 03 June 2022;
Published: 01 July 2022.

Edited by:

Sujit Datta, Princeton University, United States

Reviewed by:

Jay Nadeau, Portland State University, United States
Yuan-Nan Young, New Jersey Institute of Technology, United States

Copyright © 2022 Vachier  and Wettlaufer . This is an open-access article distributed under the terms of the Creative Commons Attribution License (CC BY). The use, distribution or reproduction in other forums is permitted, provided the original author(s) and the copyright owner(s) are credited and that the original publication in this journal is cited, in accordance with accepted academic practice. No use, distribution or reproduction is permitted which does not comply with these terms.

*Correspondence: Jérémy Vachier , jeremy.vachier@su.se, John S. Wettlaufer , john.wettlaufer@yale.edu

Disclaimer: All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors and the reviewers. Any product that may be evaluated in this article or claim that may be made by its manufacturer is not guaranteed or endorsed by the publisher.