J. Phys. Soc. Jpn. 92, 121009 (2023) [8 Pages]
SPECIAL TOPICS: Advances in the Physics of Biofluid Locomotion

Cellular Ethological Dynamics in Diorama Environments

+ Affiliations
1Research Center of Mathematics for Social Creativity, Research Institute for Electronic Science, Hokkaido University, Sapporo 001-0020, Japan2Graduate School of Life Science, Hokkaido University, Sapporo 001-0020, Japan3Faculty of Engineering, University of the Ryukyus, Nishihara, Okinawa 903-0213, Japan

How smart can cells behave under complicated environmental conditions in their natural habitat? Further, what physical mechanism brings about such smart behaviors? These two questions are closely related to each other and are to be studied in both cell biology and physics. Here, we focus on behavioral smartness and variability observed mainly in unicellular eukaryotic organisms. We review a few typical cases in which cell behavior seemed smart in a complex experimental condition and/or variable in a similar experimental condition. On each of these behaviors, a possible mechanism was considered by formulating and analyzing the model equations of cell motion. By revisiting the thought that cells might be simple machines that could repeat the same response to the same stimuli and conditions, we threw some light on the plausible mechanisms for behavioral smartness and variability in cellular movements.

©2023 The Physical Society of Japan
1. Introduction: A Scope of Cellular Ethological Dynamics
Well-organized movements of crawling amoebae

The behavior of protists (mostly unicellular eukaryotes) has often been the subject of physical research. One of the typical examples is the collective locomotion exhibited by soil amoeba (Dictyostelium discoideum) in starved conditions: rotating spiral waves emitted by thousands of amoebae that are interacting with each other by the chemical diffusion of chemotactic attractant (cAMP, cyclic adenosine monophosphate). This mechanism was modeled primarily by an excitable reaction-diffusion equation,1,2) similar to the pattern formation of rotating spirals seen in the Belousov–Zhabotinski reactions.3,4) In reality, rich patterns of rotating spiral motions were observed, and reproduced secondarily by the model equations of reaction-diffusion-advection (the advection term came from chemotactic movement) type.58) The richness was brought about by the differences in parameters and initial conditions of model equations that may be plausible in real systems. This self-organized collective motion enables the organism to form a fruiting body for reproduction.

Smart behaviors of a giant amoeba

The behavioral ability of a giant amoeba, plasmodium of Physarum polycepharum, to solve a foraging task under some complicated situations like a maze has been discovered.9,10) Two food pellets were presented after the plasmodium spread over the maze. The organism gathered at the both food pellets while forming a vein between two food pellets through the maze. The thick vein often traced nearly the shortest route between two food pellets by cutting a corner through a center-in-center trajectory. At this time, the thick shortest vein was not always unique but some thinner veins remained in sub-shortest routes. The connection pattern of veins displayed a variety.

The mechanism for the network formation of veins was modeled using differential equations that described a simplified biological rule of current reinforcement.1115) The model equations could reproduce the variety of vein networks by some differences in initial conditions and parameters.1618) This model is introduced in detail in Sect. 2.

Variability of tactic movement of swimming microalgae

In addition to amoeboid cells, another type of cell, which could migrate by cilia (microalgae, ciliate, flagellate, gamete of multi-cellular organisms, etc.), has been extensively studied well for a long time. As recent advances in biological research have discovered that flagella and cilia have the same molecular structure, researchers in this field intended to re-refer to flagella as cilia. In this study, we follow this style of referencing.

For instance, the phototactic movement of microalga Chlamydomonas has been rigorously studied well in cell biology and fluid dynamics. The phototactic sign [i.e., positive (attraction) or negative (repulsion)] depended on the light intensity illuminated and on the intracellular conditions of redox potential due to photosynthesis.1923) The movement to the light stimuli could be switched between positive (attractive) and negative (repulsive). This switching mechanism was still unclear but considered by the mechanical motion of their characteristic method of spiral swimming and its temporal modulation of ciliary beating induced by sensing of light stimulation.24,25)

A colony of microalgae like Gonium and Volvox showed phototaxis as well but controlling phototactic direction may be nontrivial because it is not simple to determine what coordinated modulation of ciliary beating from cell to cell determines the tactic direction.26,27) Computational bio-fluid dynamics was necessary to bridge this wide gap between the local response in each cell to light exposure and the tactic direction of the whole colony.27,28) It turned out that a slight quantitative difference in the local response could change the tactic direction. This implies potential variability of behavior in the same environmental conditions. A similar potential of behavioral variability was expected not only in a colony but also in a single microalga. A model of phototaxis in a single-celled alga Chlamydomonas is introduced in detail later in Sect. 4.

For swimming cells, mechanical environments are influential. The geometrical shape of the surrounding space and the presence of solid obstacles affected a swimming trajectory. Reaction to these mechanical constraints was not just simple collision but varied modulation in their swimming manner. They could change the swimming trajectory in response to fluid flow without direct contact with a solid surface. This means that they could react to each other and to a wall and an obstacle through fluid motion. Thus hydrodynamic study was essential as fluid motion mediated the interaction with mechanical environments and other micro organisms.

Toward physical elucidation of cell behavior in complex conditions

In recent years, research on cell behavior has entered a new phase due to improvements in measurement technology and computer simulation technology, as well as advances in mathematical analysis. As shown in the previous sections, we introduced some examples of behavioral smartness and variability. In general, it might be expected that cells are merely simple machines that can repeat the same response to the same conditions. However, previous studies suggested that it was reasonable to revise this conventional thought.2934) Hence, we need to re-examine the possibility that the smartness and variability of cell behavior is higher than usually assumed.

In this review, we have focused mainly on protists and introduced the state of research that approaches the multiple options of behavioral actions in experimental conditions. While introducing some concrete examples, emphasis was made on our methodological key points: (1) How to design the complex conditions that actualize potential capacity of behavioral variability (diorama environments, a keyword we proposed), and (2) how to formulate the model equations that describe behavioral variability (ethological dynamics, another keyword).35) One of the goals is to understand the behavioral adaptability of cells that may be expectedly higher than one usually assumed. Such behavioral adaptability may be the basis of information processing in multi-cellular organisms like animals, plants and fungi (proto-intelligence, a conceptual keyword, which means that it is different from but related to human intelligence).

2. A Typical Example of Smart Behavior in a Giant Amoeba, Physarum Plasmodium
Highly smart behavior in the field where many small food pellets are distributed

Single-celled eukaryote Physarum plasmodium (a giant amoeba) can find a shorter path through a maze by tracing a center-in-center trajectory at zigzag corners. Further, as a foraging behavior, the organism can construct a multi-objective transport network among spatially distributed small food pellets.10,16) As the multi-objective network among the food pellets is as well-designed as the man-made railway network of public transportation in the Tokyo region, the potential ability of single-celled organisms may be higher than one usually expected.

Such foraging behavior of slime molds is related to their habitats: the gloomy and moist environment where they are deposited with rotting wood and litter on the forest floor. The interior of decaying trees has a porous structure with large gaps, making it like a complicated labyrinth. Among them, small animals, their excrement, corpses, fungi, bacteria, decomposed organic matter etc., are scattered here and there. The environment inside the decaying tree is not homogeneous in space and not stationary in time as there will be diurnal fluctuations at a minimum. Crawling around in such an uneven and unsteady environment, they find scattered food.

Behavioral smartness exerted in complex environments

Based on careful observations of wild habitats, simulated field environments can be designed in the laboratory by focusing on certain environmental factors. For example, focusing on the complexity of the spatial arrangement of scattered food-sources, the same food pellets could be scattered in complex pattern of food-positions under constant temperature and humidity in the dark. The slime mold expanded its body according to this food distribution and found a scattered food pellet. If the simulated field environment designed in the laboratory is appropriate, the slime mold will show a potential ability of adaptive behavior to complex conditions. A simulated outdoor environment inspired by the habitat is an effective way to bring out the latent behavioral abilities of cells. We referred to these simulated environments as diorama environments.

Another way to awaken cellular potential is to use well-established intelligence tests (tasks). There are habituation and learning tests that are used in behavioral psychology and comparative cognitive (neuronal) science. Maze-solving is one of them.36,37) Because such a test method is a complicated environment designed in a laboratory, this is also referred to as a diorama environment.

Mathematical modeling for smart foraging behavior

It turned out that Physarum plasmodium can connect the separated food pellets by efficient routing of plasmodial veins and pass around many obstacles. They established a network of veins between distributed food pellets. This behavior was based on the adaptive growth of the vein diameter that worked locally in each vein: it became thicker when the flow through the vein was enough, and thinner if otherwise (hereafter we call this the current reinforcement rule).11,13)

Current reinforcement model for adaptive development of transport network to the complex conditions

The amoeboid body of Physarum plasmodium can be regarded as a network of water pipes.12,1618) Suppose the water flows in and out from food pellets as protoplasmic mass is much thicker at the food pellets and flows between them through the network of veins. A cylindrical hard pipe \(E_{ij}\) that connects two joints \(N_{i}\) and \(N_{j}\), has length \(l_{ij}\), and radius \(r_{ij}\), and flow rate \(q_{ij}\). According to Poiseulle flow, \(q_{ij} = d_{ij}\frac{p_{i}-p_{j}}{l_{ij}}\), where \(d_{ij}=\frac{\pi r_{ij}^{4}}{8\eta}\) and \(p_{i}\) is the pressure at the joint \(N_{i}\). Due to the volume conservation, \(\sum_{j\sim i} q_{ij} =0\) at any joint \(N_{i}\), except the source and sink of water flow (food pellets). Please note that the expression \(\sum_{j\sim i}\) means summation of j that are connected to \(N_{i}\). Water volume \(q_{0}\) flows in at the source and out at the sink, \(\sum_{j\sim i} q_{ij} =+ q_{0}\) and \(-q_{0}\), respectively. When p at the sink is assumed to be zero, all of the \(p_{i}\) and \(q_{ij}\) are determined.

Next, we describe the current reinforcement rule for pipe growth \(\frac{d d_{ij}}{dt}= f(|q_{ij}|)-\alpha d_{ij}\), where \(f(q_{ij})\) is the current reinforcement function (CRF) that is a monotonically increasing function of \(q_{ij}\), but the function is not recognized from experiments. We assume a power function \(f(|q_{ij}|)= |q_{ij}|^{\alpha}\) at the beginning and, later, a sigmoidal function to be more realistic. This ordinary differential equation includes the peculiar properties of the current reinforcement in Physarum plasmodium.

The network shape is different depending on the slight difference in the variable (initial thickness of each pipe) and parameters α, \(q_{0}\), and CRF as well. When the number of food pellets increases, and their distribution pattern becomes more complicated, the process of network formation from time to time seems efficient enough to be an advanced ability for survival.38)

Elucidation of a heuristic algorithm for the smart behavior embedded in the current reinforcement model

Interestingly, this heuristic mechanism of network formation of Physarum was modeled by dynamical equations of motion. The model equations were simply based on a physical equation of motion for the network flow and some addition of a biological feature of adaptation (current reinforcement rule). Mechanics-based description of smart cell behavior in complicated situations is a new focus in physics because such studies to explore an unexpected way of biological information processing in a single cell without any neurons.

Please note the necessity to simplify the real situation in such modeling. In fact, in this model, plasmodial veins were regarded as rigid pipes, like water pipes. By disregarding complex rheology, the effect of the current reinforcement rule could be examined.

Such models that describe cell behavior with a small number of macro-scopic variables (e.g., continuum mechanics) contain properties peculiar to living organisms. The properties are that the boundary conditions and parameters of the equation can change depending on the state of the system. These changes originate from active movements of organisms, from changes in the environment caused by such movements, and/or from changes in the complex physical properties of substances that constitute cells. In any case, the model is based on mechanics. In this sense, we call such a model equation the ethological dynamics of cells. The goal is to describe the mechanism of information processing by the time evolution equation so that it can be used as a heuristic algorithm for solving a problem in the different fields of science and technology.3942)

The current reinforcement model for the foraging behavior of Physarum was simple and worked well in the complicated pattern of food distribution. The model dynamics is the heuristic algorithm, which is a repetitive update of the thickness of a pipe through a specific relationship between flow through the pipe itself and the growth rate of pipe thickness (the current reinforcement function). This heuristic algorithm produces network geometries that meet the multiple requirements in complex patterns of food distribution. In this sense, behavioral smartness is induced by the complexity of the environment.

Moreover, slight differences in the variables and parameters give rise to different network shapes. The final shapes are diverse. This diversity brings unexpected variability and plasticity to the foraging behavior of slime molds. Ethological dynamics exhibited in such a complex environment often result in advanced ability in problem-solving (survival strategy). In this sense, this advanced ability to survive in a complex environment is a primitive form of intelligent organismic heuristics (proto-intelligence), distinguishing it from the intelligence of higher animals.

3. Reversibility of Peristaltic Crawling in Cells and Multicellular Organisms

Amoeboid crawling, which involves a sol–gel transformation of the cell, can also be viewed as a coupled system of local motion that is often bistable, excitable, or oscillatory in terms of dynamics. It is known that these three dynamic behaviors arise from the same equations of motion (with different parameters).43) In fact, these dynamic behaviors observed in Physarum and Dictyostelium amoebas were modeled by differential equations.44)

In the giant amoeba Physarum plasmodium, local rhythmic contraction propagates throughout the body, thereby receiving a force (reaction force) from the substrate surface to propel it forward. Such movement is called peristaltic crawling (the word peristalsis means propagation of contraction waves) and can be considered in the same way as the movement of earthworms and gastropods because all these organisms can migrate with periodic propagation of contraction waves along the body. It is known that these organisms can switch the moving directions depending on the situation, but the dynamic mechanism has not been obvious.45)

Various mechanical models of peristaltic crawling have been proposed.46) Here, we will introduce a simple model.4749) For the sake of simplicity, the body of the organism is assumed to be a one-dimensional body consisting of alternating mass points and springs (Fig. 1). The point mass has weight and only has kinetic friction (the static friction ignored) with the substrate surface. The equation of motion was \begin{equation} \xi \dot{X}_{n} = K \{(X_{n+1}-X_{n}-l_{n+\frac{1}{2}})-(X_{n}-X_{n-1}-l_{n-\frac{1}{2}})\}, \end{equation} (1) where \(X_{n}\) was the position of mass point n, ξ was the viscous friction coefficient, \(l_{n+1/2}\) was the natural length of the Hookean spring between the mass points n and \(n+1\), K was the spring constant, and \(\dot{X}\) was the time derivative of X. In this equation, an overdamping regime was assumed.


Figure 1. A model for peristaltic crawling in one-dimensional space.47) (a) Schematic diagram for multi-block model. Each block had the mass and a dynamic friction coefficient with the substrate surface and was connected to each other by Hookean springs. However, the natural length of this spring changed depending on time and space. This represented wave propagation of peristaltic motion. (b) In the non-dimensionalized model, the peristaltic wave e (wavelength λ) propagated at velocity v, shown at the bottom of the figure. The periodic rise (δ-function) of friction coefficient η was given at a certain phase of the wave and moved with the wave at velocity v, as shown in the upper panel. (c) The dependency of migration velocity on the phase of friction rise in the non-dimensionalized model. The migration velocity was averaged over one period (λ) of the peristaltic wave. The solution \(u^{(0)}\) was obtained in the unperturbed condition (the friction coefficient was constant), and the friction rise was given at any phase β. Depending on β, the migration velocity V (dotted line) changed. When \(V<0\) (the left half of the figure), \(u^{(0)}_{t}>0\) (this part of the body was moving to the right), and \(u^{(0)}_{s}<0\) (this part was contracting and denser). Conversely, when \(V>0\) (the left half of the figure), that part was moving to the left while relaxing and elongating. Please note that the part of the body that was moving to the right was always contracted while the part moving to the left was always elongated. When the brakes (the friction rise) were applied in the contracted part (moving to the right), the body migrated to the left (\(V<0\)). Contrary to this, when the body migrated to the right, the brakes were applied in the elongated part (moving to the left).

To represent active peristaltic motion, the natural length of a spring (Hookean spring) was varied periodically, and its phase propagated along the body as \begin{equation} l_{n+\frac{1}{2}} = l_{0} +a \sin (\omega t-2 \pi n/N), \end{equation} (2) where ω was angular velocity, N was the spatial period, and a and \(l_{0}\) were the constants.

These equations were rewritten as \begin{align} \xi_{n} \dot{U}_{n} &= K[U_{n+1}+U_{n-1}-2U_{n}\notag\\ &\quad-a \{\sin (\omega t-2\pi n/N)-\sin (\omega t-2\pi (n-1)/N)\}], \end{align} (3) by introducing the new variable \(U_{n}=X_{n}-n l_{0}\), where \(U_{n}\) was the displacement of mass point n from the standard position.

By introducing the reference coordinate \(s=n l_{0}\) and the dimensionless variables and parameters, \(\eta_{n}(s, t)=\xi_{n}(t)/l_{0}\) (rescaled friction constant), \(\kappa =l_{0} K\) (rescaled spring constant), \(e_{0}=a/l_{0}\), \(k=2\pi/(Nl_{0})\), and by taking the limit \(l_{0}\rightarrow 0\), the continuum equations were obtained \begin{align} \eta(s, t) \frac{\partial u}{\partial t} = \kappa \left(\frac{\partial^{2} u}{\partial s^{2}}-\frac{\partial e}{\partial s}\right), & \end{align} (4) \begin{align} e(s, t) = e_{0} \sin (\omega t-ks). & \end{align} (5)

When the friction coefficient η was constant in time and space, the net displacement in one cycle of contraction motion was zero, although local cyclic displacement could be observed. If the friction coefficient was constant and anisotropic (different in anterior-posterior directions), it moved in the direction of low friction, but in this case, it was not able to move in the opposite direction.

What mechanism enables the variability of forward and backward movement? This issue was considered next. So, it was assumed that the friction constant changed with periodic motion as the body of the organism expanded and contracted due to rhythmic contraction. For example, in Physarum plasmodium, it might change due to some biochemical and/or mechanical conditions on the substrate surface. In worms and gastropods, it was expected that the body was pressed against the substrate surface or lifted. Although the details were still unknown, it was plausible that the friction constant changed at a certain phase of rhythmic motion.

For simplicity and convenience of formulation, a rise in friction constant was introduced with δ-function at the certain phase (δ-function pulse of anchoring) as \(\eta\rightarrow \eta\{1+\epsilon p(s, t)\}\), where \(p(s, t)=\sum_{n=-\infty}^{\infty}2\pi\delta \{\omega t -k(s-B-n\lambda)-\alpha\}=\sum_{n=-\infty}^{\infty}2\pi\delta \{\omega t -(ks-\beta-2n\pi)-\alpha\}\) (where \(\beta=kB\)), ϵ was the small parameter for perturbation expansions, β was the phase distance of the δ-function pulse measured from a peak of peristaltic wave \(e(s, t)\), λ was the wave length, and α was constant. The governing equation of motion to be solved was \begin{equation} \{1+\epsilon\,p(s, t)\} \frac{\partial u}{\partial t} = D_{0}\left\{\frac{\partial^{2} u}{\partial s^{2}}+v_{0} \cos (\omega t-k s)\right\}, \end{equation} (6) where \(D_{0}=\kappa/n_{0}\) and \(v_{0}=ke_{0}\).

By assuming the perturbation expansion of solution \(u=u^{(0)}+\epsilon\,u^{(0)}+\cdots\), the governing equations with the 0-th and 1st order of ϵ were obtained as \begin{align} \frac{\partial u^{(0)}}{\partial t} = D_{0} \left\{\frac{\partial^{2} u^{(0)}}{\partial s^{2}}+v_{0}\cos(\omega t-ks)\right\}, & \end{align} (7) \begin{align} \frac{\partial u^{(1)}}{\partial t} = D_{0} \frac{\partial^{2} u^{(1)}}{\partial s^{2}}-p(s, t)\frac{\partial u^{(0)}}{\partial t}, & \end{align} (8) and solved as \begin{equation} \left\langle \frac{\partial u^{(1)}}{\partial t} \right\rangle_{\lambda} {}= - \omega A \sin \beta, \end{equation} (9) where \(\langle\bullet \rangle_{\lambda}=\frac{1}{\lambda}\int^{\lambda}_{0}(\bullet)\,ds\). The migration distance averaged over the wavelength λ is \(V=\epsilon\langle u_{t}\rangle^{(1)}_{\lambda}=-\epsilon\omega A\sin \beta =-\epsilon e_{0}\frac{\omega}{k}\cos \alpha\sin \beta\).

When the frictional anchoring took place in a compressed part of the body, the direction of migration was opposite to the direction of the peristaltic wave. Conversely, when the anchoring was in a stretched part of the body, the migration direction was the same as the peristaltic wave. The conclusion was that (1) even if the peristaltic wave propagated in the same direction, the phase difference of anchoring switched the migration direction.

From the symmetry of the equation of motion, (2) even if the phase of anchoring was the same, the migration direction was reversed when the propagation direction of the peristaltic wave was reversed. Furthermore, (3) even if both the propagation direction of the peristaltic wave and the phase of anchoring were the same, the migration direction reversed when the frictional pulse switched from anchoring (increase of friction constant) to releasing (decrease of frictional constant).

As described above, three factors contribute to the reversal of the direction of movement. Both directions of movement are inherent in the mechanism of peristaltic crawling. This physical nature gives the potential variability of behavior to the crawling cells and organisms. This insight is to be stressed, when thinking about the evolutionary development of behavioral flexibility and diversity.

Therefore, it is important to efficiently control the friction constant accompanied by the local periodic motion. If there were some mobile appendages on the body, efficient control would be possible, and the legs of animals like lugworms, centipedes, millipedes, caterpillars, and insects might be considered as the appendages that acquired such functionality.

In the future, if we can formulate the meta-clonal waves of ciliary beating in a similar framework of mechanical modeling, we will be able to describe the crawling motion and the swimming motion in a unified manner.

4. Switching of Phototaxis in a Swimming Microalgae Chlamydomonas

A microalga Chlamydomonas shows phototaxis. Its sign depends on the light intensity: positive (attractive) to weak light and negative (repulsive) to overpowering light. When it receives light, photosynthesis takes place and the redox potential within the cell changes. This redox potential controls the sign of phototaxis. There is an intracellular mechanism to switch the sign of phototaxis according to the environment.

Conversely, even under the same experimental conditions, there are cell-to-cell differences. Even under conditions where most cells show positive phototaxis, a few cells show negative or neutral motility. Individual cells that eventually exhibit positive phototaxis may initially be temporarily away from light. Although most individual cells show positive taxis, each cell takes a different option of behavioral responses.

The mechanism of such behavioral variability and diversity has been investigated with a mechanical model. Chlamydomonas has a spheroid-like body, with a single eye spot in the middle of the body that senses light from the surroundings (Fig. 2). At the tip of the body (one end of the long axis of the spheroid), two cilia move like the arms in human breaststroke, so the body moves forward. At this time, the cell body rotates around a certain axis (an axis with some angular difference from the long axis of the spheroid). This allows Chlamydomonas to trace a spiral trajectory of swimming. Inevitably, the eye spot senses light while rotating 360 degrees.


Figure 2. Model of spiral swimming in Chlamydomonas.25) The coordinate system \(\vec{a}\), \(\vec{b}\), and \(\vec{c}\) was attached to the body of the organism and represented the direction of the body orientation. The axis \(\vec{c}\) was the anterior-posterior axis of the body, and it was assumed that the organism moved at a constant speed \(v_{0}\) along the \(\vec{c}\) direction. Note that the eye spot was directed toward the direction \((\vec{a}+\vec{b})/2\), and the cilium closer to this eye spot was called cis-cilium (or cis-flagellum), and the other one was called trans-cilium (or trans-flagellum). Chlamydomonas was assumed to be a rigid body and to rotate around the \(\vec{\omega}\)-axis. The body orientation of Chlamydomonas was described in the coordinate system \((x, y, z)\) fixed in the laboratory.

When the eye spot receives light (turns in the direction from which the light comes), the rhodopsin photoreceptor mechanism in the eye spot is activated, and with a short delay time, the beating patterns of cilia are changed temporarily. Then the body axis of Chlamydomonas tilts and the swimming trajectory changes. The intensity of the light received affects the temporal modulation of ciliary beating. In addition, even under the same experimental conditions, the temporal modulation of ciliary beating and the lag time from photoreception varies considerably among individuals. As described above, some variability of phototactic response is observed in Chlamydomonas.

To consider the phototactic variability, various simple models of spiral swimming have been proposed.25,26) The Chlamydomonas body was replaced by a rigid spheroid. A Cartesian coordinate system with the long axis of the spheroid as the c-axis (with the other two axes as the a- and b-axes) was introduced to always determine the anterior-posterior, left-right, and dorsoventral axes of the body. The spheroidal Chlamydomonas was assumed to move always forward at a constant speed \(v_{0}\) in the positive direction of the c-axis while rotating around the ω-axis, which has some angular differences with the a-, b-, c-axes. The eye spot was assumed to be oriented to the direction of \((\vec{e}_{a}+\vec{e}_{b})/2\). However, here, \(\vec{e}_{a}\) and \(\vec{e}_{b}\) were unit vector in the positive direction of the a- and b-axes, respectively.

Assuming that the ω-axis tilted temporarily when the eye spot received light. In this case of photoresponse, there were three parameters: the tilting direction and magnitude of ω-axis, and the lag time between the photo-reception and the tilting. The photo-reception \(I(t)\) (photo-reception function) was given by the inner product for the direction from which the light came (parallel rays) and the direction to which the eye spot faced, \(I(t)=I_{0}(-\vec{e}_{\textit{light}}\cdot\vec{e}_{\textit{eyespot}}+1)/2\).

The rotation vector \(\vec{\omega}\) (rotation around the ω-axis) was in general decomposed into three rotation components around three axes, \(\vec{\omega}=\omega_{a} (t)\vec{a} +\omega_{b} (t)\vec{b} +\omega_{c} (t)\vec{c}\), where \(\omega_{a} (t) =\omega_{a}^{(0)}-p(t)\), \(\omega_{b} (t) =0\), and \(\omega_{c} (t) =-\omega_{c}^{(0)}\) [1/s], in a specific model.25) Here \(p(t)\) was photo-induced tilting of \(\vec{\omega}\), given by \(p(t)=\gamma_{0}\frac{d I(t-\tau_{0})}{dt}\), where \(\tau_{0}\) was the lag time, and \(\gamma_{0}\) was −1 or +1 for the two opposite directions of tilting (this came from the experimental results). At last, all three parameters were of the tilting magnitude \(I_{0}\), the tilting direction \(\gamma_{0}\), and the lag time of tilting \(\tau_{0}\).

The translational and rotational motion of spheroid was described in 3-d space by the space-fixed Cartesian coordinate \((\vec{x},\vec{y},\vec{z})\) as \begin{align} \frac{d\theta_{1}}{dt} = - \frac{\cos \theta_{3}}{\sin \theta_{2}}(\omega_{a}^{(0)}-p(t)), \end{align} (10) \begin{align} \frac{d\theta_{2}}{dt} = \cos \theta_{3} (\omega_{a}^{(0)}-p(t)), \end{align} (11) \begin{align} \frac{d\theta_{3}}{dt} = - \omega_{c}^{(0)}+\cos \theta_{3} \frac{\cos \theta_{2}}{\sin \theta_{2}} (\omega_{a}^{(0)}-p(t)), \end{align} (12) where \(\theta_{1}\), \(\theta_{2}\), and \(\theta_{3}\) were Euler angles. These were the governing equations for the model.

After the study on the dynamic behavior of the model, the following was found. Depending on the parameters of \(\tau_{0}\) and \(\gamma_{0}\), these were the finding. Sign of the chemotaxis was switched. In addition, the movement of temporarily moving away from the light and finally approaching to the light source was reproduced. The swimming trajectory did not always approach towards (recede from) light monotonically. Between the photo-induced transient modulation of ciliary beating and the final expression (phenotype) of chemotaxis, there was an inherent gap that gave rise to potential behavioral diversity, and the final destination changed depending on external and internal conditions.

What was interesting from a mathematical point of view was that there existed a swimming trajectory in which the eye point always faced at the constant angle with respect to the direction in which the light ray came (steady-state solutions). There were two steady-state solutions for each positive and negative phototaxis. In a set of parameters of the model, the linear stability of these steady-state solutions was determined. For the steady-state solutions of positive or negative phototaxis, one was stable and the other was unstable. No bistable regions were present. Both of positive and negative phototaxis lay in the dynamical mechanism for spiral swimming. Such latent behavioral variability can be a source of diverse behaviors even in very similar conditions.

5. Transference of Protist Behaviors to Multi-cellular Organisms

Rotating spiral motion observed in Dictyostelium was observed in other multicellular organisms.50) It may be seen in the contraction of the heart.5159) When the heart was in a disordered state of tachycardia, a rotating spiral wave of contraction was observed, and later contraction pattern of the heart could be a critical stage of fibrillation. In this case, the rotating spiral wave is a nuisance to be removed, and its effective removal method has been physically studied. Additionally, in connection with the fertilization of Xenopus oocytes, a rotating spiral wave of calcium ion was observed on the cell surface.60) The rotating spiral wave is common in both protists and vertebrates as a dynamical behavior of excitable cells.

In the formation of the transport network observed in the Physarum plasmodium, the network shape was similar to that of public transportation in the Tokyo region in the sense of multi-objective optimization: cost of total length, fault-tolerance of global connectivity against accidental disconnection of local edge, and efficiency of transportation between two sites. This similarity could be based on the similar mechanism of current reinforcement. Each edge was variable adaptively to the protoplasmic and traffic flow in Physarum and public transportation network, respectively.16) It is interesting that a similar mechanism worked in these two substantially different systems although it might be coincidental. A similar mechanism to the current reinforcement is found in all of three groups of multicellular eukaryotes: fungi,61) plants,6264) and animals.65)

In addition to amoeboid tactic movement, ciliary tactic movement was observed in Chlamydomonas and was found in many other species of protist. This ciliary motion was characterized commonly by the spiral trajectory of swimming. Sperms in multicellular organisms swam using flagella (its molecular structure is the same as cilium) and showed a similar spiral trajectory to an ovum or multiple ova by sensing chemo-attractant. In the sense of spiral swimming with sensing, methods of tactic behavior by cilia could be compared in protist and multicellular organisms.

In fertilization, so many sperms swim together in a semi-closed internal space of a female body or an open space of natural habitat with an inevitable disturbance.6670) At the time, they not only competed but also cooperated with each other. This interaction among ciliary spiral swimmers is interesting from a hydrodynamic point of view. Collective motions of red tide algae and Euglena could be a research subject to be studied by both hydrodynamics and cell biology.7173) In a colony of microalgae like Gonium and Volvox, phototactic behavior was established through hydrodynamic interaction between many cilia, while similar hydrodynamic interaction could be expected in 2d sheet of ciliary cells on the surface of the mammalian airway and the surface of the internal cavity in sponges. Further, comparative studies on ciliary interaction in protists and multi-cellular organisms are needed.

6. Summary

The outdoor environment inhabited by protists is never uniform and flat and changes over time, considering their body size and range of migration. In this sense, they live in a sufficiently complex environment. Simultaneously, there is also the pressure of predator-prey relationships. As in some examples already mentioned in this review, protist behavior is smarter and more variable than one might expect, even though they are single-celled organisms. Consequently, their behavior shows diversity and advanced ability in problem-solving. These abilities have commonalities with cell-level behaviors in multi-cellular organisms and appear to be inherited from protists to multi-cellular organisms. Therefore, the elucidation of the diversity of cell behavior and the mechanism of problem-solving ability is useful for a basic understanding of the ability of information processing in multi-cellular organisms.

In terms of basic science, this research squarely tackles the classic question of how an organism's ability for information processing arises from the physical motion of matter. A cell's response to a single stimulus is generally simple and transient. However, due to the complexity of the environment and the active migration into new environments, cells continue to receive various stimuli in a complicated temporal sequence. Such sequential stimuli have complex effects on multiple time-scale dynamics within the cell body, resulting in diverse and smart behavioral outputs. Therefore, smart behavior can only be demonstrated in the corresponding complex environment.

To summarize, we shall conclude this review by quoting a short excerpt from the project website35) that includes the three keywords.

Intelligence broadly describes an ability to adapt to the environment. In this sense, single-celled organisms like protists (eukaryotic unicellular organisms) have a prototype of intelligence, or rather they can demonstrate skillful behavior in complex field environments due to their sophisticated evolution over hundreds of millions of years. This behavioral ability seems to be inherited as “single-cellular” behavior in multicellular organisms (sperm motility during fertilization, cell motility in the internal environment, etc.).

In this Research Area, we define “proto-intelligence” as the fundamental adaptability to the environment that single-celled organisms potentially possess. We name such artificial conditions as “diorama environments” where organisms can show their potential proto-intelligence. Diorama environments may mimic the complexity of a habitat but in a setup designed for testing proto-intelligence. For example, one such instance is that of an amoeboid organism of slime mold, which displays the ability to find the shortest path in a maze of diorama environments.

Because the mechanisms of proto-intelligence can often be formulated using coupled kinetic equations of cell motion and the environment, such environment-coupled mechanics will be thoroughly applied. We will challenge and advance the algorithms (heuristics) of proto-intelligence.

Acknowledgment

All authors would like to thank the members of the research project MEXT Kakenhi No. 21A402 Advanced mechanics of cell behavior shapes formal algorithm of protozoan smartness awoken in diorama conditions (or in short, Ethological Dynamics in Diorama Environments) for helpful and suggestive discussions. This work was supported by the grant-in-aid MEXT Kakenhi (Nos. 21H05303 and 21H05310).


References

  • 1 J. J. Tyson and J. D. Murray, Development 106, 421 (1989). 10.1242/dev.106.3.421 CrossrefGoogle Scholar
  • 2 J. Halloy, J. Lauzeral, and A. Goldbeter, Biophys. Chem. 72, 9 (1998). 10.1016/S0301-4622(98)00119-7 CrossrefGoogle Scholar
  • 3 K. I. Agladze and V. I. Krinsky, Nature 296, 424 (1982). 10.1038/296424a0 CrossrefGoogle Scholar
  • 4 S. C. Müller, T. Plesser, and B. Hess, Science 230, 661 (1985). 10.1126/science.230.4726.661 CrossrefGoogle Scholar
  • 5 E. F. Keller and L. A. Segel, J. Theor. Biol. 30, 225 (1971). 10.1016/0022-5193(71)90050-6 CrossrefGoogle Scholar
  • 6 T. Höfer and P. K. Maini, Phys. Rev. E 56, 2074 (1997). 10.1103/PhysRevE.56.2074 CrossrefGoogle Scholar
  • 7 E. Vidal-Henriquez, V. Zykov, E. Bodenschatz, and A. Gholami, Chaos 27, 103110 (2017). 10.1063/1.4986153 CrossrefGoogle Scholar
  • 8 P. Schaap, Philos. Trans. R. Soc. B 376, 20190756 (2021). 10.1098/rstb.2019.0756 CrossrefGoogle Scholar
  • 9 T. Nakagaki, H. Yamada, and A. To’th, Nature 407, 470 (2000). 10.1038/35035159 CrossrefGoogle Scholar
  • 10 T. Nakagaki, H. Yamada, and A. To’th, Biophys. Chem. 92, 47 (2001). 10.1016/S0301-4622(01)00179-X CrossrefGoogle Scholar
  • 11 T. Nakagaki, H. Yamada, and T. Ueda, Biophys. Chem. 84, 195 (2000). 10.1016/S0301-4622(00)00108-3 CrossrefGoogle Scholar
  • 12 A. Tero, R. Kobayashi, and T. Nakagaki, J. Theor. Biol. 244, 553 (2007). 10.1016/j.jtbi.2006.07.015 CrossrefGoogle Scholar
  • 13 T. Nakagaki and R. D. Guy, Soft Mater. 4, 57 (2008). 10.1039/B706317M CrossrefGoogle Scholar
  • 14 R. D. Guy, T. Nakagaki, and G. B. Wright, Phys. Rev. E 84, 016310 (2011). 10.1103/PhysRevE.84.016310 CrossrefGoogle Scholar
  • 15 D. Akita, T. Nakagaki, D. Schenz, S. Kuroda, K. Sato, and K. Ueda, Dev. Growth Differ. 59, 465 (2017). 10.1111/dgd.12384 CrossrefGoogle Scholar
  • 16 A. Tero, S. Takagi, T. Saigusa, K. Ito, D. P. Bebber, M. D. Fricker, K. Yumiki, R. Kobayashi, and T. Nakagaki, Science 327, 439 (2010). 10.1126/science.1177894 CrossrefGoogle Scholar
  • 17 Q. Ma, A. Johansson, A. Tero, T. Nakagaki, and D. J. T. Sumpter, J. R. Soc. Interface 10, 20120864 (2013). 10.1098/rsif.2012.0864 CrossrefGoogle Scholar
  • 18 D. Akita, I. Kunita, M. D. Fricker, S. Kuroda, K. Sato, and T. Nakagaki, J. Phys. D 50, 024001 (2017). 10.1088/1361-6463/50/2/024001 CrossrefGoogle Scholar
  • 19 M. E. H. Feinleib and G. M. Curry, Physiol. Plant. 25, 346 (1971). 10.1111/j.1399-3054.1971.tb01453.x CrossrefGoogle Scholar
  • 20 T. Kondo, C. H. Johnson, and J. W. Hastings, Plant Physiol. 95, 197 (1991). 10.1104/pp.95.1.197 CrossrefGoogle Scholar
  • 21 N. Isogai, R. Kamiya, and K. Yoshimura, Zool. Sci. 17, 1261 (2000). 10.2108/zsj.17.1261 CrossrefGoogle Scholar
  • 22 T. Takahashi and M. Watanabe, FEBS Lett. 336, 516 (1993). 10.1016/0014-5793(93)80867-T CrossrefGoogle Scholar
  • 23 K. Wakabayashi, Y. Misawa, S. Mochiji, and R. Kamiya, Proc. Natl. Acad. Sci. U.S.A. 108, 11280 (2011). 10.1073/pnas.1100592108 CrossrefGoogle Scholar
  • 24 K. C. Leptos, M. Chioccioli, S. Furlan, A. I. Pesci, and R. E. Goldstein, bioRxiv, version 2 (2018) [DOI: 10.1101/254714]; Google ScholarK. C. Leptos, M. Chioccioli, S. Furlan, A. I. Pesci, and R. E. Goldstein, Phys. Rev. E 107, 014404 (2023). 10.1103/PhysRevE.107.014404 CrossrefGoogle Scholar
  • 25 M. Nakajima, K. Iizuka, N. Ueki, A. Isu, K. Yoshimura, T. Nakagaki, T. Hisabori, K. Sato, and K. Wakabayashi, bioRxiv (2020) [DOI: 10.1101/2020.12.06.414052]. Google Scholar
  • 26 H. de Maleprade, F. Moisy, T. Ishikawa, and R. E. Goldstein, Phys. Rev. E 101, 022416 (2020). 10.1103/PhysRevE.101.022416 CrossrefGoogle Scholar
  • 27 R. E. Goldstein, Annu. Rev. Fluid Mech. 47, 343 (2015). 10.1146/annurev-fluid-010313-141426 CrossrefGoogle Scholar
  • 28 R. E. Goldstein, J. Fluid Mech. 807, 1 (2016). 10.1017/jfm.2016.586 CrossrefGoogle Scholar
  • 29 H. S. Jennings, Behaviour of the Lower Organisms (Columbia University Press, New York, 1906). CrossrefGoogle Scholar
  • 30 E. M. Eisenstein, Aneural Organisms in Neurobioloby (Plenum, New York, 1972). Google Scholar
  • 31 W. C. Corning and R. Von Burg, Protozoa, in Invertebrate Learning: Protozoans through Annelids, ed. W. C. Corning and J. A. Dyal (Plenum, New York, 1973) Vol. 1, p. 49. 10.1007/978-1-4684-3006-6_2 CrossrefGoogle Scholar
  • 32 D. Bray, Wetware: A Computer in Every Living Cell (Yale University Press, New Haven, CT, 2011). Google Scholar
  • 33 R. Brette, eNeuro 8, ENEURO.0018-21.2021 (2021). 10.1523/ENEURO.0018-21.2021 CrossrefGoogle Scholar
  • 34 S. J. Gershman, P. Balbi, C. R. Gallistel, and J. Gunawardena, eLife 10, e61907 (2021). 10.7554/eLife.61907 CrossrefGoogle Scholar
  • 35 Web [https://diorama-ethology.jp/eng/]. Google Scholar
  • 36 M. Hainworth, Experiments in Animal Behavior (Macmillan, London, 1967). Google Scholar
  • 37 M. Jitsumori and S. Nakajima, Gakushu no Shinri (Saiensusha, Tokyo, 2023). Google Scholar
  • 38 K. Ito, D. Sumpter, and T. Nakagaki, Nonlinear Theory Its Appl., IEICE 1, 26 (2010). 10.1587/nolta.1.26 CrossrefGoogle Scholar
  • 39 A. Tero, R. Kobayashi, and T. Nakagaki, Physica A 363, 115 (2006). 10.1016/j.physa.2006.01.053 CrossrefGoogle Scholar
  • 40 A. Tero, T. Nakagaki, K. Toyabe, K. Yumiki, and R. Kobayashi, Int. J. Unconv. Comput. 6, 109 (2010). Google Scholar
  • 41 T. Nakagaki, A. Tero, R. Kobayashi, I. Onishi, and T. Miyaji, New Generation Comput. 27, 57 (2008). 10.1007/s00354-008-0054-8 CrossrefGoogle Scholar
  • 42 C. Gao, C. Liu, D. Schenz, X. Li, Z. Zhang, M. Jusup, Z. Wang, M. Beekman, and T. Nakagaki, Phys. Life Rev. 29, 1 (2019); 10.1016/j.plrev.2018.05.002 Crossref;, Google ScholarC. Gao, C. Liu, D. Schenz, X. Li, Z. Zhang, M. Jusup, Z. Wang, M. Beekman, and T. Nakagaki, Phys. Life Rev. 29, 51 (2019). 10.1016/j.plrev.2019.07.002 CrossrefGoogle Scholar
  • 43 J. Murray, Mathematical Biology I -An Introduction- (Springer, Berlin, 2002). Google Scholar
  • 44 D. Imoto, N. Saito, A. Nakajima, G. Honda, M. Ishida, T. Sugita, S. Ishihara, K. Katagiri, C. Okimura, Y. Iwadate, and S. Sawai, PLOS Comput. Biol. 17, e1009237 (2021). 10.1371/journal.pcbi.1009237 CrossrefGoogle Scholar
  • 45 J. Gray, Animal Locomotion (George Weidenfeld and Nicolson Ltd., London, 1968). Google Scholar
  • 46 M. Alexander, Principles of Animal Locomotion (Princeton University Press, Princeton, NJ, 2002). CrossrefGoogle Scholar
  • 47 Y. Tanaka, K. Ito, T. Nakagaki, and R. Kobayashi, J. R. Soc. Interface 9, 222 (2012). 10.1098/rsif.2011.0339 CrossrefGoogle Scholar
  • 48 S. Kuroda, I. Kunita, Y. Tanaka, A. Ishiguro, R. Kobayashi, and T. Nakagaki, J. R. Soc. Interface 11, 20140205 (2014). 10.1098/rsif.2014.0205 CrossrefGoogle Scholar
  • 49 S. Kuroda, Yo. Tanaka, and T. Nakagaki, Keisoku to Seigyo (Measurement and Control) 54, 248 (2015) [in Japanese]. Google Scholar
  • 50 S. Sawai, P. Thomason, and E. Cox, Nature 433, 323 (2005). 10.1038/nature03228 CrossrefGoogle Scholar
  • 51 M. A. Allessie, F. I. M. Bonke, and F. J. G. Schopman, Circ. Res. 41, 9 (1977). 10.1161/01.RES.41.1.9 CrossrefGoogle Scholar
  • 52 J. P. Keener, SIAM J. Appl. Math. 46, 1039 (1986). 10.1137/0146062 CrossrefGoogle Scholar
  • 53 A. T. Winfree, J. Theor. Biol. 138, 353 (1989). 10.1016/S0022-5193(89)80200-0 CrossrefGoogle Scholar
  • 54 J. Davidenko, A. Pertsov, R. Salomonsz, W. Baxter, and J. Jalife, Nature 355, 349 (1992). 10.1038/355349a0 CrossrefGoogle Scholar
  • 55 O. Steinbock, V. Zykov, and S. C. Mueller, Nature 366, 322 (1993). 10.1038/366322a0 CrossrefGoogle Scholar
  • 56 A. T. Winfree, Science 266, 1003 (1994). 10.1126/science.7973648 CrossrefGoogle Scholar
  • 57 J. Jalife, Annu. Rev. Physiol. 62, 25 (2000). 10.1146/annurev.physiol.62.1.25 CrossrefGoogle Scholar
  • 58 A. Karma, Annu. Rev. Condens. Matter Phys. 4, 313 (2013). 10.1146/annurev-conmatphys-020911-125112 CrossrefGoogle Scholar
  • 59 R. Majumder, S. Hussaini, V. S. Zykov, S. Luther, and E. Bodenschatz, PLOS Comput. Biol. 17, e1009476 (2021). 10.1371/journal.pcbi.1009476 CrossrefGoogle Scholar
  • 60 J. Keener and J. Sneyd, Mathematical Physiology (Springer, Berlin, 1998). CrossrefGoogle Scholar
  • 61 L. Heaton, B. Obara, V. Grau, N. Jones, T. Nakagaki, L. Boddy, and M. D. Fricker, Fungal Biol. Rev. 26, 12 (2012). 10.1016/j.fbr.2012.02.001 CrossrefGoogle Scholar
  • 62 H. Fujita and A. Mochizuki, J. Theor. Biol. 241, 541 (2006). 10.1016/j.jtbi.2005.12.016 CrossrefGoogle Scholar
  • 63 C. Feller, E. Farcot, and C. Mazza, PLOS ONE 10, e0118238 (2015). 10.1371/journal.pone.0118238 CrossrefGoogle Scholar
  • 64 J. Haskovec, H. Jönsson, L. M. Kreusser, and P. Markowich, Proc. R. Soc. A 475, 20190015 (2019). 10.1098/rspa.2019.0015 CrossrefGoogle Scholar
  • 65 T. Miura, Fukuoka Acta Med. 107, 161 (2016).10.15017/1787037 CrossrefGoogle Scholar
  • 66 K. Shiba, S. A. Baba, T. Inoue, and M. Yoshida, Proc. Natl. Acad. Sci. U.S.A. 105, 19312 (2008). 10.1073/pnas.0808580105 CrossrefGoogle Scholar
  • 67 M. Abdelgalil, Y. Aboelkassem, and H. Taha, Phys. Rev. E 106, L062401 (2022). 10.1103/PhysRevE.106.L062401 CrossrefGoogle Scholar
  • 68 S. Phuyal, S. S. Suarez, and C. K. Tung, Front. Cell Dev. Biol. 10, 961623 (2022). 10.3389/fcell.2022.961623 CrossrefGoogle Scholar
  • 69 J. F. Jikeli, L. Alvarez, B. M. Friedrich, L. G. Wilson, R. Pascal, R. Colin, M. Pichlo, A. Rennhack, C. Brenker, and U. B. Kaupp, Nat. Commun. 6, 7985 (2015). 10.1038/ncomms8985 CrossrefGoogle Scholar
  • 70 S. Lange and B. M. Friedrich, PLOS Comput. Biol. 17, e1008826 (2021). 10.1371/journal.pcbi.1008826 CrossrefGoogle Scholar
  • 71 T. Shikata, S. Kitatsuji, and K. Yuasa, Front. Cell Dev. Biol. 11, 1134227 (2023). 10.3389/fcell.2023.1134227 CrossrefGoogle Scholar
  • 72 N. J. Suematsu, H. Yamashita, and M. Iima, Front. Cell Dev. Biol. 11, 1134002 (2023). 10.3389/fcell.2023.1134002 CrossrefGoogle Scholar
  • 73 T. J. Pedley and J. O. Kessler, Annu. Rev. Fluid Mech. 24, 313 (1992). 10.1146/annurev.fl.24.010192.001525 CrossrefGoogle Scholar

Author Biographies


Yukinori Nishigami was born in Osaka prefecture, Japan in 1984. He obtained his B.Sc. (2008), M.Sc. (2010), and D.Sc. (2013) degrees from the University of Hyogo. After working as a postdoctoral fellow at Kyoto University, he has been an assistant professor at Hokkaido University since 2018. He has been working on the behaviors and locomotion mechanisms of protists.

Itsuki Kunita was born in Hokkaido Prefecture, Japan in 1982. He earned his Dr. of Sys. Info. Sci. from Future University Hakodate in 2011. He was a postdoctoral fellow at Future University Hakodate (2011–2014), Hokkaido University (2014–2015), and Kumamoto University (2015–2016), and an assistant Professor at University of the Ryukyus (2016–2022). He is an associate professor at University of the Ryukyus. His research focuses on the behaviors and locomotion mechanisms of single-celled organisms.

Katsuhiko Sato is an associate professor at Research Institute for Electronic Science, Hokkaido University from 2014. He got his Ph.D. in Physics (1999) from Kyoto University on the subject of polymer melt. He has worked on non-equilibrium physics, soft matter physics and biophysics as a COE at Kyoto University (2000–2002), a postdoctoral research fellow at the University of Tokyo (2002– 2007), an assistant professor at Tohoku University (2007–2011), and a scientific researcher at RIKEN in Kobe (2011–2014). He is currently working on shear banding, phototaxis in ciliates, and collective cell migration.

Toshiyuki Nakagaki was born in Aichi prefecture, Japan in 1963. He obtained his B.Sc. (1987), M.Sc. (1989) degrees from Hokkaido University. After working for a pharmaceutical company, he completed his doctorate (Ph.D. in 1997) at Nagoya University, while working as a part-time lecturer at a correspondence high school. He has been a professor at Hokkaido University since 2013. He has been working on the physical ethology, mainly in protists.