Cellular responses to pheromone in yeast can range from gene expression to morphological and physiological changes. While signaling pathways are well studied, the cell fate decision-making during cellular polar growth is still unclear. Quantifying these cellular behaviors and revealing the underlying physical mechanism remain a significant challenge. Here, we employed a hidden Markov chain model to quantify the dynamics of cellular morphological systems based on our experimentally observed time series. The resulting statistics generated a stability landscape for state attractors. By quantifying rotational fluxes as the non-equilibrium driving force that tends to disrupt the current attractor state, the dynamical origin of non-equilibrium phase transition from four cell morphological fates to a single dominant fate was identified. We revealed that higher chemical voltage differences induced by a high dose of pheromone resulted in higher chemical currents, which will trigger a greater net input and, thus, more degrees of the detailed balance breaking. By quantifying the thermodynamic cost of maintaining morphological state stability, we demonstrated that the flux-related entropy production rate provides a thermodynamic origin for the phase transition in non-equilibrium morphologies. Furthermore, we confirmed that the time irreversibility in time series provides a practical way to predict the non-equilibrium phase transition.

Inducing haploid Saccharomyces cerevisiae with pheromone, a sex hormone that promotes mating causes yeast cells to cease asexual reproduction in favor of heterosexual cell fusion, resulting in diploids.1–3 Due to the inability of yeast cells to move like flagellated bacteria, they can only grow in the direction of a high pheromone concentration after inhibiting their own cell cycle when stimulated by pheromone continuously.4–6 Fus3p, an essential MAP kinase that shuttled back and forth to the nuclear membrane, was responsible for all signal transduction in the cellular response.7–15 It mediates polar growth through two distinct molecular pathways: P1 pathway (P1: Fus3 → Far1 → Cdc24 → Cdc42 → Bni1 → polar growth)7,16–18 and P2 pathway (P2: Fus3 → Bni1 → polar growth).8,19 Different levels of negative feedback regulation were induced by high- and low-dose pheromone in the biological processes underlying cell response (Fig. 1). Once pheromone concentrations surpass a certain threshold, the yeast cells activate stronger signaling inhibition, including the indirect inhibition of Fus3 by Ste12.20,21 While the signaling pathways by which cells respond to pheromone have received significant attention in biology,1,22–24 little is known about the cell fate decision-making that occurs during cellular polar growth.25 Quantifying the dynamics and thermodynamics of the cell fate decision-making reflected in such biological processes as cell growth and morphology remains a significant challenge.

FIG. 1.

A mechanistic network wiring upon different concentrations of pheromone. P1 and P2 represent the two signaling pathways for the polar growth; I1–I3 represent the inhibitory effects of the negative feedback regulation; Fus3_inner and Fus3_outer, respectively, represent Fus3 in the nucleus and the cytoplasm.

FIG. 1.

A mechanistic network wiring upon different concentrations of pheromone. P1 and P2 represent the two signaling pathways for the polar growth; I1–I3 represent the inhibitory effects of the negative feedback regulation; Fus3_inner and Fus3_outer, respectively, represent Fus3 in the nucleus and the cytoplasm.

Close modal

Although the interacting potential energy landscapes can quantify the global behavior and the local dynamics in equilibrium, these dynamical rules are not directly applicable to the living biological systems.26,27 A living biological system itself is not in equilibrium due to its continuous exchange of information, matter or energy with the external environment.26,28,29 Indeed, in non-equilibrium systems, the driving forces for the dynamics can be quantified in terms of the underlying landscape's gradient and the curl probability fluxes.31–34 Such a dynamical complex system is intrinsically open, breaking the thermodynamic equilibrium in which detailed balance can be maintained.33,34 This underlying landscape, defined by the non-equilibrium potential U, is inextricably linked to the steady-state probability distribution, which quantifies the probability of hills and valleys forming in specific states. Meanwhile, the probabilistic steady-state fluxes are forces with curl or rotating properties that can be used to precisely measure the degree of the detailed balance breaking.26,27 Such non-zero curl fluxes, by breaking the detailed balance, can have a significant effect on both the energy dissipation in biological systems and the irreversibility of dynamical processes.35,36 This is because the curl flux is the dynamical origin of entropy production or dissipation, and the entropy production rate quantifies the thermodynamic cost of maintaining a non-equilibrium system's steady state.34,37 As a result, the system's detailed balance is disrupted, resulting in time reversal asymmetry and determining the direction of time.

Along with describing dynamics, the mathematical theory of Markov processes has increasingly been used to develop a thermodynamic theory describing the transformation of energy between all forms in recent years.38–41 Markov systems with discrete state variables and continuous time enable thermodynamics to be related to dynamics' theory of potential energy landscapes.39,41 In general, the cell morphology characterized by area or perimeter is the result of cumulative and irreversible cell growth. If the Markov dynamics can capture these processes, then one can link the nonequilibrium dynamics and nonequilibrium thermodynamics to reveal the underlying biological process.26,34,42–44

Here, microscopy was used in this study to monitor the living state of single-celled yeast cultured in an isothermal microfluidic device in real time. We observed four distinct cell morphological fates using a specific form (Hn) to characterize cell morphology.25 Markov processes are used to describe changes in a cellular morphological system. By quantifying the landscape of cell morphological fates at various concentrations, we found that between 2.0 and 3.0 μM, there were distinct tendencies for the phase transformations in cell morphology toward a dominant single fate. To elucidate the non-equilibrium dynamics of cellular fate decision-making, the probabilistic fluxes of the cell morphological landscape were quantified. Moreover, the quantification of non-equilibrium thermodynamics confirmed that the induction of high-dose pheromones costs yeast cells more energy, supporting the molecular mechanism of greater negative feedback at high doses. The mutual validation of the biological experiments and physical theories, as well as the dynamics and thermodynamics, confirmed the viability and precision of using a non-equilibrium statistical physical theory to quantify living biological systems. This attempt to establish the physical link between molecular events and cellular characteristics across scales has significant implications for future research into the non-equilibrium nature of cell fate decision-making in living biological systems.

A molecular network containing both the positive feedback to activate Fus3 and the negative feedback (I1–I3) to directly or indirectly inhibit Fus3 can, in principle, cause yeast cells to reach a bistable state over time (Fig. 1). To determine whether the yeast cells can achieve a non-equilibrium steady state in response to the external stimuli, the expressions of FUS3-GFP in single yeast cells were monitored over time [Fig. 2(a) and Figs. S1 and S2].25 After approximately 600 min of cultivation, the gene expression levels of Fus3 outside the nucleus indicate that the yeast cells have entered a steady state, in which the fluorescence intensity of Fus3 fluctuates around a certain value. Meanwhile, the population landscape of the Fus3 (cytoplasmic and nuclear fluorescent signals) reveals the emergence of the bistable state in the non-equilibrium steady-state phase [Figs. 2(b) and S3].25 

FIG. 2.

Quantification of the pheromone-induced cellular fate decision-making in yeast.25 (a) Trajectories of fluorescence intensity of Fus3 at 0.7 μM inside and outside the nucleus (only a portion shown). The dashed gray line at 600 min was used to approximate the time node at which all cell fluorescence trajectories had entered a non-equilibrium steady state. (b) The population landscape of the Fus3 fluorescence intensity inside and outside the nucleus of the yeast cells in the stationary phase at 0.7 μM pheromone. (0,0) and (1) represent two steady states of the Fus3 gene expressions. (c) A simple diagram of cell shape with circle filling pattern. The image on the left depicts a yeast cell cultured for 1440 min in a medium containing 1.0 μM pheromone; the red line indicates the contour of the cell; the image on the right is the filling model for the image on the left; R 1 R 3 are the radii of the circle. (d) Real-time trajectory of the cell morphology ( H n) at 0.7 μM. The red horizontal line is for approximating the position of the four cell morphologies. (e) Normalized probability distribution of cell morphology in response to different pheromone concentrations. Normalized method: P x d x = 1; the sample sizes at steady state for each pheromone concentration are as follows: 0.7 μM was equivalent to 21 335 cells, 0.8 μM to 18 408 cells, 1.0 μM to 23 041 cells, 2.0 μM to 36 886 cells, and 3.0 μM to 18 276 cells. (f) Yeast cells cultured with varying pheromone doses were observed under a microscope. The left panel at each pheromone concentration depicts the cell morphology without pheromone addition; the right panel depicts the cell morphology after 1380 min of incubation in YNB medium supplemented with pheromone; the intracellular green fluorescence is the light excited by Fus3p-GFP under a 488 nm laser.

FIG. 2.

Quantification of the pheromone-induced cellular fate decision-making in yeast.25 (a) Trajectories of fluorescence intensity of Fus3 at 0.7 μM inside and outside the nucleus (only a portion shown). The dashed gray line at 600 min was used to approximate the time node at which all cell fluorescence trajectories had entered a non-equilibrium steady state. (b) The population landscape of the Fus3 fluorescence intensity inside and outside the nucleus of the yeast cells in the stationary phase at 0.7 μM pheromone. (0,0) and (1) represent two steady states of the Fus3 gene expressions. (c) A simple diagram of cell shape with circle filling pattern. The image on the left depicts a yeast cell cultured for 1440 min in a medium containing 1.0 μM pheromone; the red line indicates the contour of the cell; the image on the right is the filling model for the image on the left; R 1 R 3 are the radii of the circle. (d) Real-time trajectory of the cell morphology ( H n) at 0.7 μM. The red horizontal line is for approximating the position of the four cell morphologies. (e) Normalized probability distribution of cell morphology in response to different pheromone concentrations. Normalized method: P x d x = 1; the sample sizes at steady state for each pheromone concentration are as follows: 0.7 μM was equivalent to 21 335 cells, 0.8 μM to 18 408 cells, 1.0 μM to 23 041 cells, 2.0 μM to 36 886 cells, and 3.0 μM to 18 276 cells. (f) Yeast cells cultured with varying pheromone doses were observed under a microscope. The left panel at each pheromone concentration depicts the cell morphology without pheromone addition; the right panel depicts the cell morphology after 1380 min of incubation in YNB medium supplemented with pheromone; the intracellular green fluorescence is the light excited by Fus3p-GFP under a 488 nm laser.

Close modal

Given that Fus3-activated Far1 can indicate the polarity direction of the cell growth, P1 pathway makes cell growth more directional than P2 pathway does.7,16,19,45 As Fus3 switches between its two stable states, the relative weight of the two pathways of polar growth (P1 and P2) alters, thereby influencing the process of cell polar growth. When observing yeast cell growth under a microscope in real time, we noticed two alternating patterns in the polar growth of yeast cells: first, more lateral than longitudinal growth, followed by more longitudinal than lateral growth. Therefore, the quantification of cell growth should take into account of the growth in length and width of the cell's various parts, rather than relying solely on area or perimeter. We segmented yeast cells using a circular fill pattern [Fig. 2(c)] and employed a value similar to the harmonic mean (Hn) to characterize cell morphology.25 Hn equals the sum of the reciprocal radii of the filled circles multiplied by the number of filled circles, i.e., H n = n ( 1 R 1 + 1 R 2 + + 1 R n ). The cell morphological trajectories reveal a refined cell growth process in which Hn increases as the cell grows longer along the long axis and decreases as the cell grows wider along the short axis [Fig. 2(d) and Fig. S4].25 Around 600 min, the cell morphology achieves a specific steady state that corresponds to a growth mode in which different morphologies can alternate.

By calculating the distribution of Hn at various concentrations, four distinct cell morphological fates in yeast cells at steady state (data after 600 min) were identified [Fig. 2(e)].25 These four morphological fates can, of course, be observed at any time during the non-equilibrium steady-state period due to the abundance of statistical data volumes. To facilitate the comparison of cell morphology at different pheromone concentrations, the morphology distribution was normalized, i.e., P x d x = 1. Notably, as pheromone concentrations increased, the four morphological fates observed at low pheromone doses (0.7–2.0 μM) tend to merge into a single dominant fate at high pheromone dose (3 μM). The proportion of “State2” significantly increases, while the proportion of other states decreases correspondingly. That is, at doses near 3 μM, the tendency of the phase transition in cell morphology appears to emerge.

In fact, this tendency of phase transition in cell morphology is observable during microfluidic yeast cell cultivation. The size of cell morphology was significantly greater at concentrations of 0.7–2.0 μM pheromone than that at a concentration of 3.0 μM [Fig. 2(f)].25 From the perspective of the mechanical network wiring, the morphological change at 3.0 μM is caused by the difference in yeast cells' negative feedback regulation. Once the pheromone concentration surpasses a certain threshold, Ste12 will enter the cytoplasm from the nucleus in the yeast cell to activate Msg5, thereby inhibiting the activity and the expression of Fus3.20,21 Thus, the high-dose pheromone-induced inhibitors (I1–I3) inhibit the expression level and the activity of Fus3 more effectively than the low-dose inhibitors (I1 and I2), leading to variable degrees of polar growth (Fig. 1). This stronger negative feedback regulation induced by this high-dose pheromone caused a phase transition in the cell's morphological system, resulting in a change in the scales used to measure the four morphological fates. Consequently, the coordinates of the cell morphology statistics chart deviated from the coordinates of other concentrations at 3.0 μM.

To quantify the cell morphological fates in this non-equilibrium steady state, we simulated the dynamics between cell morphologies using a four-state Markov model. The transition between the four morphologies is a Markov process in discrete state space, and the equation determining the dynamics of the probability distribution P is analogous to the Fokker–Planck equation in the continuous representation,46,47 which can be represented as
d p i d t = j T j i P j j T i j P i .
(1.1)
Here, P i is the probability of state i. T is the transition rate matrix that represents the transition rate in cell morphology from one state to another. For example, T i j is the transition rate from state i to state j. Obviously, the transition between states i and j is balanced by increasing or decreasing the probability of their corresponding states, thus the master equation implies probability conservation d i P i / d t = 0. Similarly, Eq. (1.1) can be written as
d P / d t = M T P ,
(1.2)
with P is the probability vector for all discrete cell morphologies and M is the transition probability matrix or the transition rate matrix among different states obtained from the experimental data. M i j = T i j for i j, M i j = 1 j T i j for i = j.
In our system, the transition of cell morphology in response to pheromone stimulation occurs in a non-equilibrium steady state. It is evident from d P / d t = M T P that the transition rate matrix M controls the evolution of the probabilistic dynamics of the cell morphology. To quantify the non-equilibrium dynamics of this biological system, we can decompose the driving force for cell morphological transition into two components: a time-reversible component D that holds a detailed balance (potential landscape) and a time-irreversible component C that breaks this detailed balance (flux landscape)
M = C + D .
(1.3)
While C and D can be given by
C i j = { max ( T i j P i s s T j i P j s s , 0 ) / P i s s , i j ( 1 ) j C i j , i = j ,
(1.4)
D i j = { min ( T i j P i s s , T j i P j s s ) / P i s s , i j ( 1 ) j C i j , i = j .
(1.5)
The superscript “ s s” in this case stands for the steady state.

The nonequilibrium potential function or landscape ( U) can be defined as the logarithm of the probability distribution at steady state [Fig. 2(d)], that is U = ln P s s .26,30,48 It is analogous to the Boltzmann distribution law P exp β U that links the potential energy to equilibrium probabilities.41,49 If U = ln P s s, we have D i j = min T i j , T j i e Δ U i j, here Δ U i j stands for the difference or gradient in U. Therefore, the time-reversible part ( D) of the driving force corresponds to the gradient of the non-equilibrium steady-state potential ( U). Biologically, such a potential landscape quantifies the weights of the four cell morphological states in the non-equilibrium stable state. Meanwhile, the nonequilibrium weight landscape, which is determined by the steady-state probabilities in stochastic dynamics, attracts the cell morphological states into stable landscape basins with lower potential or greater probability.

The other irreversible part ( C) of the driving force is the curl flux. For the steady state, the flux between any two states i and j can be expressed as
F i j s s = T j i P j s s T i j P i s s .
(1.6)
Since the statistically steady state is expected to emerge over a long time limit, the probability's time derivative becomes zero, that is d P i / d t = 0, causing the probability to cease to vary with time. If the F i j s s = 0, this system satisfying the detailed balance will achieve equilibrium state without generating any net local flux.39,50,51 However, d P i / d t = 0 simply indicates that the sum of all fluxes entering and exiting state i is zero ( j F i j s s = 0), whereas the individual flux between i and j state ( F i j s s) does not have to be zero.26,30 We can easily prove that D i j P i s s = D j i P j s s and if C i j P i s s > 0, C j i P j s s = 0.39,52 According to the definition of the net flux in the non-equilibrium steady state, Eq. (1.6) can be rewritten as
F i j = C j i P j s s C i j P i s s = F j i .
(1.7)
If F i j is not equal to zero, the flux of this cell morphology system ( J i j = C i j P i s s, i j) can be decomposed into a finite number of closed flux loops ( r i j), with the sum of their values equal to the detailed balance breaking part, i.e., J i j = k = 1 N r i j k (see the Appendix for a detailed mathematical discussion and derivation of flux loops).37,39,41,52–54

As a result, the curl flux decreased gradually at 0.7–2.0 μM, but significantly increased at 3.0 μM [Fig. 3(a) and Tables S1–S5]. It appears that the sudden increase in the cycle flux or the degree of nonequilibrium is consistent with the distinct morphological changes or trends of phase transition near the 3.0 μM pheromone concentration. This can be understood as follows: while the potential landscape tends to attract the system to the attractor state, the cycle flux being rotational tends to destabilize the current attractor states and this gives the opportunities of repopulating the new attractor states. In other words, the nonequilibrium characterized by the cycle flux can provide the dynamical origin of nonequilibrium phase transition.

FIG. 3.

The non-equilibrium dynamics and thermodynamics of cell fate decision-making at different pheromone concentrations. (a) The average net fluxes vs pheromone concentrations. (b) The average absolute value of the non-zero chemical current vs pheromone concentrations. (c) The average absolute value of the non-zero chemical voltage vs pheromone concentrations. (d) EPR vs pheromone concentrations. (e) Time irreversibility vs pheromone concentrations. Δ C 3 represents the difference between the forward and backward three-point correlation functions of cell morphological trajectories. (F) The one-dimensional underlying potential energy landscape U ( U = ln P) at 0.7 and 3 μM. ①–④ stand for the four morphological states.

FIG. 3.

The non-equilibrium dynamics and thermodynamics of cell fate decision-making at different pheromone concentrations. (a) The average net fluxes vs pheromone concentrations. (b) The average absolute value of the non-zero chemical current vs pheromone concentrations. (c) The average absolute value of the non-zero chemical voltage vs pheromone concentrations. (d) EPR vs pheromone concentrations. (e) Time irreversibility vs pheromone concentrations. Δ C 3 represents the difference between the forward and backward three-point correlation functions of cell morphological trajectories. (F) The one-dimensional underlying potential energy landscape U ( U = ln P) at 0.7 and 3 μM. ①–④ stand for the four morphological states.

Close modal
To elucidate the precise mechanism by which the detailed balance is broken and the flux becomes non-zero, we investigated the thermodynamic origin of the rotational flux in the cellular morphological landscape. The chemical flux between distinct cell morphological fates can be expressed as
J i j = T j i C j T i j C i ,
(2.1)
where C i , C j are the “concentrations” of the State ( i ) and State ( j ) in the cell morphological fates, respectively; the T i j and T j i are the transition probabilities to be taken from master Eq. (1.1). Due to the fact that the concentrations of these states vary over time, the C can be written as
d C / d t = j J i j .
(2.2)
By choosing P i = C i / j C j , the j C j can be considered as a constant, so Eq. (2.1) can be rewritten as
J i j = T j i P j T i j P i ,
(2.3)
which is similar to Eq. (1.1). Indeed, the J i j is the generalized thermodynamic flux of the reaction State ( i ) State ( j ). The conjugate generalized thermodynamic driving force ( A i j) for generating the thermodynamic flux can be used to define the generalized chemical potential,27,34,39,55–57 i.e.,
A i j = ln T j i P j / T i j P i .
(2.4)
We obtain this thermodynamic force by dividing the affinity with the constant k B T, where k B denotes the Boltzmann constant and T is the cell's absolute temperature in culture.

To better explain the thermodynamic origin of the cell morphological system, the chemical flux ( J i j ) can be viewed as the circuit's current ( I) and the chemical potential ( A i j) as the circuit's voltage ( V). The results demonstrate that both the chemical currents and chemical voltages against pheromone concentrations have similar behaviors as the curl steady state probability flux against the pheromone concentrations corresponding to the states of various cell morphologies, i.e., the significant increase in the concentration at 3.0 μM [Figs. 3(b) and 3(c)]. Meanwhile, we determined the Pearson correlation coefficients of net flux and chemical current, net flux and chemical voltage to be 0.9936 and 0.9351, respectively [Figs. 4(a) and 4(b)]. Overall, the chemical voltage and chemical current form a dual for determining the underlying biochemical kinetics and the nonequilibrium thermodynamics.

FIG. 4.

Correlations between different nonequilibrium physical quantities in the cell morphological system. (a) The correlation between the net flux and the chemical current. “PCC” is “Pearson correlation coefficient”; the circles represent the experimental data for the net flux and the chemical current at five concentrations of the pheromones; the line is the fitted curve. (b) The correlation between the net flux and the chemical voltage. (c) The correlation between the net flux and the EPR. (d) The correlation between the net flux and the time irreversibility.

FIG. 4.

Correlations between different nonequilibrium physical quantities in the cell morphological system. (a) The correlation between the net flux and the chemical current. “PCC” is “Pearson correlation coefficient”; the circles represent the experimental data for the net flux and the chemical current at five concentrations of the pheromones; the line is the fitted curve. (b) The correlation between the net flux and the chemical voltage. (c) The correlation between the net flux and the EPR. (d) The correlation between the net flux and the time irreversibility.

Close modal
Comparing to low dose, high dose of pheromone induced stronger negative feedback regulation, resulting in substantial phase transition in morphological landscape (Fig. 1). Physically, the high-dose landscape has a deeper basin or attractor than the low-dose landscape, which may result in a greater cost of energy for cells to switch from State2 to other states [Fig. 3(f)]. To characterize the nonequilibrium thermodynamics, one needs to identify the thermodynamic cost asseverated with the process. Such cost can be represented by the entropy production rate (EPR). This cell morphological system's entropy can generally be defined as39,58
S = i P i ln P i .
(3.1)
Unlike in equilibrium thermodynamics, the entropy production in non-equilibrium thermodynamics is not solely due to the dissipation of free energy during equilibrium relaxation. This is because in non-equilibrium thermodynamics, the total entropy production rate includes both relaxation dissipation rate and the housekeeping contribution to maintain the detailed balance breaking steady state, which is from the non-zero curl flux.33,34,39,58,59 In other words, the change in total entropy of this cell morphological system with respect to time ( S ̇ tot) can be attributed to the entropy change of the system ( S ̇) and the entropy change or heat dissipation rate from the surrounding environment ( S ̇ E), which can be defined as follows:
S ̇ = S ̇ tot S ̇ E .
(3.2)
By employing bilinear expressions for the generalized flux ( J i j ) and the generalized force ( A i j), the change in the total entropy with respect to time, denoted by S ̇ tot, can be expressed as
S ̇ tot = i , j T i j P i T j i P j ln T i j P i T j i P j .
(3.3)
The heat dissipation rate from environment, denoted by S ̇ E, can be expressed as
S ̇ E = i , j T i j P i T j i P j ln T i j T j i .
(3.4)
In the non-equilibrium steady state, the entropy change of the non-stationary spontaneous relaxation part does not need be considered (it is equal to zero), and thus, the total entropy production rate of the system can be calculated using the portion of the housekeeping heat required to maintain a steady state, i.e.,
S ̇ tot s s = i , j T i j P i s s T j i P j s s ln T i j P i s s T j i P j s s = i , j T i j P i s s T j i P j s s ln T i j T j i .
(3.5)
Consequently, the EPR can be regarded as the sum of the products of chemical potential and the net chemical flux, analogous to how electric power can be understood as the product of voltage and current.

As a result, the value of EPR decreased gradually between 0.7 and 2.0 μM and increased significantly between 2.0 and 3.0 μM [Fig. 3(d)]. At 2.0–3.0 μM, it appears that the significant change in the entropy production rate can provide the necessary thermodynamic cost for the phase transitions of the cell morphology. By calculating the correlation between probabilistic net flux and EPR at various concentrations, we determined their Pearson correlation coefficient to be 0.9954 [Fig. 4(c)]. This shows the strong correlation between the flux and the entropy production rate, which relates the dynamical origin and thermodynamic origin of the nonequilibrium morphological phase transitions. In addition, this confirms the prediction that cells with shorter polar growth consume more energy than cells with longer polar growth when negative feedback regulation is stronger.

According to the fluctuation theorem, a system becomes irreversible when its entropy production is nonzero.60,61 As the irreversibility increases, the energy dissipation emerges in the cellular morphological fate decision processes. When the non-conservative forces are present, the net flux breaks the detailed balance, resulting in the emergence of time-reversal asymmetry for this non-equilibrium system.59,62

The irreversibility can be quantified by higher-order correlation functions, and we calculated the time reversal in cell morphology ( H) using a three-point correlation function. It is the difference between forward and backward in time three-point correlation functions, i.e.,
Δ C 3 = t 1 0 d t 1 t 2 0 d t 2 ( C 3 ( t 1 , t 2 ) C 3 ( t 1 , t 2 ) ) 2 ,
(4.1)
with C 3 t 1 , t 2is denoted by
C 3 ( t 1 , t 2 ) = δ H ( τ ) δ H ( τ + t 1 ) δ H ( τ + t 1 + t 2 ) / δ H 2 3 / 2 ,
(4.2)
where δ H t denotes the cell's morphological trajectory fluctuation, i.e., δ H t = H t H.

As can be seen from the results, the value of the averaged Δ C 3 decreased gradually between 0.7 and 2.0 μM and increased significantly between 2.0 and 3.0 μM, which is consistent with the probability net flux change trend with pheromone concentration [Fig. 3(e)]. At 0.7–2.0 μM, the little drop or near-constant averaged Δ C 3 suggests that the difference between the temporal forward and backward three-point correlations is small. Meanwhile, the significant increase in the averaged Δ C 3 at 2.0–3.0 μM indicates the presence of a significant net flux, which breaks the detailed balance to a greater extent. We determined the Pearson correlation coefficients of the net flux and the averaged Δ C 3 to be 0.9700 [Fig. 4(d)]. Thus, the curl flux serves as a quantitative indicator of the degree to which the detailed balances are broken, and the time is irreversible. This shows that the tendency of nonequilibrium phase transition can also be characterized by the significant changes in the time irreversibility. Since the time irreversibility can be quantified by the high order correlations directly from the observational time series, this provides a practical way of predicting the onset of nonequilibrium phase transition even when the underlying mechanism is unknown.

The induction of pheromone caused yeast cells to undergo significant responses ranging from gene expressions to cell morphological and physiological changes. Considering that the molecular mechanism of cell polar growth involves the direction of polarity, we chose a characterization order parameter (Hn) that can account for the growth of different parts in yeast. The significant advantage of this order parameter is that it is particularly sensitive to morphological changes at various locations of the cell, allowing differential non-directional normal growth and directional polar growth (i.e., lateral growth and longitudinal growth) to be characterized in real time. Unlike conventional quantitative methods such as area and perimeter, this method of cellular characterization breaks down the notion that the cellular morphology can only accumulate irreversibly over time.

Four distinct cell morphological fates were distinguished by fitting cell morphological trajectories in non-equilibrium steady state over time using a hidden Markov chain model. The statistics of the observed time series can give rise to the landscape quantifying the stability of the state attractors. To quantify the underlying dynamics of the cell fate decision-making, we decomposed the driving force for cell morphological transition into the potential landscape and the flux landscape. Although the asymmetric flux created by the four cell morphological system states is not unique, it can be represented by the three basic loop or cycle fluxes. To quantify the level of global detailed balance breaking, we used the average of these three loops fluxes to indicate the morphological system's degree of non-equilibrium. The results demonstrated that the average net flux around a pheromone concentration of 3.0 μM also undergoes a significant phase transition [Fig. 3(a)].

The probability flux associated with the steady-state chemical flux in the biochemical kinetic system is frequently caused by the energy input from ATP hydrolysis. To visualize the microscopic chemical fluxes in the cell morphological systems, the four cell morphologies can be viewed as reactants or products in the biochemical reactions catalyzed by ATP and growth-related proteins such as Fus3p, i.e., State ( i ) ATP State ( j ). Given that the Hn-defined cell morphology is sensitive to both width and length, the ATP-driven reactions can be subdivided into two biochemical reactions characterized by their relative growth capacities, namely, State ( i ) k width > length State ( j ) and State ( j ) k width < lengt h State ( i ), where the k width > length and k width < length represent the capacity of lateral and longitudinal growth, respectively. Physically, the net flux between different morphological states is directly proportional to the difference between their transition probabilities. Thus, the high net flux of 3 μM can be explained, from a biological standpoint, by the large difference in the concentration of the morphological states, which is a result of different growth capacities [Fig. 3(a)].

To determine whether biochemical reactions result in significant thermodynamic changes, we quantified the chemical flux and chemical potential in the cell-morphological system [Figs. 3(b) and 3(c)]. It is known in physics that if a system undergoes a non-equilibrium phase transition, its energy-entropy balance and free energy will be significantly altered.63–65 With the current biological technology, it is still challenging of accurately measuring the energy required for polar growth in a single cell. To quantify the thermodynamics of cell fate decision-making, we characterized the switching of cell morphology fate as a generalized biochemical reaction. We reveal that higher chemical voltages induced by high dose of pheromone resulted in higher chemical currents, which will trigger a greater net input and, thus, more degrees of the detailed balance breaking, leading to higher probabilistic fluxes and the thermodynamic costs associated with them.

Meanwhile, the thermodynamic cost was quantified by the entropy production rate (EPR) of the cell morphological system. The significant increase in EPR at a pheromone dose of 3.0 μM confirmed the peculiar behavior that the cells with shorter polar growth consume more energy than the cells with longer polar growth. Biologically, the yeast induced by a high dose pheromone simultaneously executes positive feedback regulation to promote growth and negative feedback regulation to inhibit growth. This is equivalent to ride a bicycle with the brakes engaged, so the thermodynamic cost should be higher. Furthermore, the time irreversibility characterized by the difference in forward and backward in time cross correlations directly from the experimentally observed time series provides a practical way of quantifying the nonequilibrium and predicting the nonequilibrium morphological phase transition. In summary, we applied non-equilibrium statistical physics to a living biological system in an attempt to reveal the biological phase transition of polar growth induced by high-low dose pheromone. Quantifying the non-equilibrium dynamics and thermodynamics involved in cell fate decision-making sheds new light on the molecular mechanisms underlying pheromone-induced cellular responses.

In this section, we provide a statement regarding the source of data used in this study. Our study utilized the same original data as our previously published work.25 However, in contrast to the previous article that focused on investigating cellular fate decisions and response to signaling molecule concentrations using kinetics and energy landscape analysis, the current study takes a different perspective with a focus on establishing an understanding of non-equilibrium processes through considerations of dynamics and thermodynamics. We infer system' physical quantities and their correlations through such analysis. Specifically, in this study, we explored the same experimental data from a novel physics-oriented perspective, building upon the previous biological research to provide new insight into the field.

Saccharomyces cerevisiae S288C (ATCC 201388: MATa his3Δ1 leu2Δ0 met15Δ0 ura3Δ0) is the yeast strain used in this experiment.66, FUS3-GFP, whose C-terminus was fused with GFP as the reporter protein, was chosen as the research object from the YeastGFP database.67,68 We recovered the strain by incubating the FUS3-GFP frozen strain overnight in YPD liquid medium. Then, we isolated single colonies on YPD agar plates by inoculating them. Before observing the yeast cultured on the microfluidic plate under a microscope in real time, we selected a single colony and grew it overnight (approximately 14 h) in 5 ml YNB. The incubator's temperature was 30 °C, and its rotational speed was 250 rpm. Following is the formula for the culture medium: YPD liquid medium [Yeast extract (10 g/l), Peptone (20 g/l), and Dextrose (20 g/l)], YPD agar plate [Yeast extract (10 g/l), Peptone (20 g/l), Dextrose (20 g/l), and Agar (10 g/l)], and YNB liquid medium [Yeast Nitrogen Base Without Amino Acids (6.8 g/l), Dextrose (5 g/l), and Uracil (76 mg/l), 50 × MEM Amino Acids (20 ml/l)].

We selected pheromone, an alpha factor peptide hormone with a molecular weight of 1683.98, as the stimulus source (GenScript).69–71 On a clean bench, the weighed pheromone was dissolved in YNB liquid medium to produce a solution containing 1000 μM pheromone. Then, the microfluidic culture media containing different concentrations of pheromones were created by diluting the pheromone solution with YNB liquid medium to the optimal concentration for the experiments.

Using an inverted fluorescence microscope (Ti-E, Nikon) with an automated stage and focus and a high NA oil-immersion objective (1.45 NA, 100×), the fluorescence values of single cells were determined. The fluorescence signals were collected by a cooled EM-CCD camera using a 488 nm laser with an output power of 30 mW (only 10% of the laser beam entering the microscope objective) (897U, Andor). All images were captured utilizing both bright field and fluorescent field imaging. All bright-field and fluorescence-field images were acquired with Nikon software. Combining manual and automated analysis with the custom Matlab codes, we performed the data analysis. Multiple trajectories were acquired using time-lapse microscopy, in which fluorescent images were captured periodically and recorded every 10 min.

To investigate the underlying mechanism of the bimodality, we collected real-time fluorescence intensity trajectories [Fig. 2(a) and Figs. S1 and S2), which demonstrate that the yeast response reaches a steady state after approximately 600 min. This steady state indicates that neither the intensity of fluorescence inside nor outside the yeast cell nucleus increases or decreases significantly. The inner and outer fluorescence intensity histograms were derived from the steady-state fluorescence trajectories [Fig. 2(b) and Fig. S3]. The Hn in the yeast characterizes the cell morphology characteristics of yeast. The hidden Markov chain model (HMM) can be used to provide quantitative analysis of all cell morphology trajectories.

Using a hidden Markov model, it was possible to differentiate the cell state from the trajectory. A global maximum likelihood estimate of HMM parameters was performed on each individual time trace. Multiple initial parameters were chosen at random to initiate the iterative HMM analysis and guarantee the convergence to the global minimum. At the end of each iteration, the Baum–Welch algorithm was employed to recalculate the parameters. In each iteration, a steady-state condition was enforced on the re-estimated parameters. HMM can be applied directly to both high-dimensional and two-dimensional data. Therefore, the inner and outer fluorescence data can be trained by the HMM, and the results indicate that yeast fluorescence has a bimodal distribution. Consequently, the cell morphology trajectory data can be trained using HMM, and it is shown that the yeast cell shape has a distribution of four steady-state peaks.

S.L., Q.L., and E.W. acknowledge the support of the National Natural Science Foundation of China with Grant No. 21721003.

The authors have no conflicts to disclose.

Since this study did not involve humans or animals, ethical approval was not required.

Sheng Li: Conceptualization (equal); Data curation (lead); Formal analysis (equal); Investigation (lead); Methodology (equal); Resources (lead); Software (equal); Validation (equal); Visualization (equal); Writing – original draft (lead); Writing – review & editing (equal). Qiong Liu: Conceptualization (equal); Data curation (equal); Formal analysis (equal); Investigation (equal); Methodology (equal); Resources (equal); Software (equal); Supervision (equal); Validation (equal); Visualization (equal). Erkang Wang: Conceptualization (equal); Formal analysis (equal); Funding acquisition (lead); Investigation (equal); Methodology (equal); Project administration (lead); Resources (equal); Supervision (equal); Validation (equal); Visualization (equal). Jin Wang: Conceptualization (lead); Data curation (equal); Formal analysis (lead); Investigation (equal); Methodology (lead); Project administration (equal); Resources (equal); Software (lead); Supervision (lead); Validation (lead); Visualization (lead); Writing – original draft (equal); Writing – review & editing (lead).

The data that support the findings of this study are available from the corresponding author upon reasonable request.

1.
M. Z.
Bao
,
M. A.
Schwartz
,
G. T.
Cantin
,
J. R.
Yates
III
, and
H. D.
Madhani
, “
Pheromone-dependent destruction of the Tec1 transcription factor is required for MAP kinase signaling specificity in yeast
,”
Cell
119
,
991
1000
(
2004
).
2.
M. V.
Metodiev
,
D.
Matheos
,
M. D.
Rose
, and
D. E.
Stone
, “
Regulation of MAPK function by direct interaction with the mating-specific G alpha in yeast
,”
Science
296
,
1483
1486
(
2002
).
3.
T.
Leeuw
et al, “
Pheromone response in yeast: Association of Bem1p with proteins of the MAP kinase cascade and actin
,”
Science
270
,
1210
1213
(
1995
).
4.
V.
Sourjik
and
N. S.
Wingreen
, “
Responding to chemical gradients: Bacterial chemotaxis
,”
Curr. Opin. Cell Biol.
24
,
262
268
(
2012
).
5.
R. M.
Macnab
and
D. E.
Koshland
, “
The gradient-sensing mechanism in bacterial chemotaxis
,”
Proc. Natl. Acad. Sci. U. S. A.
69
,
2509
2512
(
1972
).
6.
L.
Bardwell
, “
A walk-through of the yeast mating pheromone response pathway
,”
Peptides
25
,
1465
1476
(
2004
).
7.
A. C.
Butty
,
P. M.
Pryciak
,
L. S.
Huang
,
I.
Herskowitz
, and
M.
Peter
, “
The role of Far1p in linking the heterotrimeric G protein to polarity establishment proteins during yeast mating
,”
Science
282
,
1511
1516
(
1998
).
8.
D.
Matheos
,
M.
Metodiev
,
E.
Muller
,
D.
Stone
, and
M. D.
Rose
, “
Pheromone-induced polarization is dependent on the Fus3p MAPK acting through the formin Bni1p
,”
J. Cell Biol.
165
,
99
109
(
2004
).
9.
M.
Peter
,
A.
Gartner
,
J.
Horecka
,
G.
Ammerer
, and
I.
Herskowitz
, “
FAR1 links the signal transduction pathway to the cell cycle machinery in yeast
,”
Cell
73
,
747
760
(
1993
).
10.
E. A.
Elion
,
B.
Satterberg
, and
J. E.
Kranz
, “
FUS3 phosphorylates multiple components of the mating signal transduction cascade: Evidence for STE12 and FAR1
,”
Mol. Biol. Cell
4
,
495
510
(
1993
).
11.
R. L.
Roberts
and
G. R.
Fink
, “
Elements of a single MAP kinase cascade in Saccharomyces cerevisiae mediate two developmental programs in the same cell type: Mating and invasive growth
,”
Genes Dev.
8
,
2974
2985
(
1994
).
12.
I.
Herskowitz
, “
MAP kinase pathways in yeast: For mating and more
,”
Cell
80
,
187
197
(
1995
).
13.
E. A.
Elion
,
J. A.
Brill
, and
G. R.
Fink
, “
FUS3 represses CLN1 and CLN2 and in concert with KSS1 promotes signal transduction
,”
Proc. Natl. Acad. Sci. U. S. A.
88
,
9392
9396
(
1991
).
14.
E.
Blackwell
et al, “
Effect of the pheromone-responsive Gα and phosphatase proteins of Saccharomyces cerevisiae on the subcellular localization of the Fus3 mitogen-activated protein kinase
,”
Mol. Cell. Biol.
23
,
1135
1150
(
2003
).
15.
F.
van Drogen
,
V. M.
Stucke
,
G.
Jorritsma
, and
M.
Peter
, “
MAP kinase dynamics in response to pheromones in budding yeast
,”
Nat. Cell Biol.
3
,
1051
1059
(
2001
).
16.
A.
Nern
and
R. A.
Arkowitz
, “
A Cdc24p-Far1p-Gβγ protein complex required for yeast orientation during mating
,”
J. Cell Biol.
144
,
1187
1202
(
1999
).
17.
A.
Nern
and
R. A.
Arkowitz
, “
Nucleocytoplasmic shuttling of the Cdc42p exchange factor Cdc24p
,”
J. Cell Biol.
148
,
1115
1122
(
2000
).
18.
Y.
Shimada
,
M.-P.
Gulli
, and
M.
Peter
, “
Nuclear sequestration of the exchange factor Cdc24 by Far1 regulates cell polarity during yeast mating
,”
Nat. Cell Biol.
2
,
117
124
(
2000
).
19.
E. H.
Chen
,
Cell Fusion: Overviews and Methods
(
Springer
,
2008
).
20.
A.
Anders
,
B.
Ghosh
,
T.
Glatter
, and
V.
Sourjik
, “
Design of a MAPK signalling cascade balances energetic cost versus accuracy of information transmission
,”
Nat. Commun.
11
,
3494
(
2020
).
21.
M.
Sugiyama
,
Y.
Kaneko
, and
S.
Harashima
, “
Yeast protein phosphatases Ptp2p and Msg5p are involved in G1–S transition, CLN2 transcription, and vacuole morphogenesis
,”
Arch. Microbiol.
191
,
721
733
(
2009
).
22.
M.
Good
,
G.
Tang
,
J.
Singleton
,
A.
Remenyi
, and
W. A.
Lim
, “
The Ste5 scaffold directs mating signaling by catalytically unlocking the Fus3 MAP kinase for activation
,”
Cell
136
,
1085
1097
(
2009
).
23.
A.
Doncic
et al, “
Compartmentalization of a bistable switch enables memory to cross a feedback-driven transition
,”
Cell
160
,
1182
1195
(
2015
).
24.
Y.
Wang
and
H. G.
Dohlman
, “
Pheromone signaling mechanisms in yeast: A prototypical sex machine
,”
Science
306
,
1508
1509
(
2004
).
25.
S.
Li
,
Q.
Liu
,
E.
Wang
, and
J.
Wang
, “
Global quantitative understanding of nonequilibrium cell fate decision making in response to pheromone
,” bioRxiv (
2023
).
26.
J.
Wang
,
L.
Xu
, and
E.
Wang
, “
Potential landscape and flux framework of nonequilibrium networks: Robustness, dissipation, and coherence of biochemical oscillations
,”
Proc. Natl. Acad. Sci. U. S. A.
105
,
12271
12276
(
2008
).
27.
S.
Lapidus
,
B.
Han
, and
J.
Wang
, “
Intrinsic noise, dissipation cost, and robustness of cellular networks: The underlying energy landscape of MAPK signal transduction
,”
Proc. Natl. Acad. Sci. U. S. A.
105
,
6039
6044
(
2008
).
28.
H.
Qian
, “
Entropy demystified: The ‘thermo’-dynamics of stochastically fluctuating systems
,”
Methods Enzymol.
467
,
111
134
(
2009
).
29.
G.
Nicolis
and
I.
Prigogine
,
Self-Organization in Nonequilibrium Systems: From Dissipative Structures to Order Through Fluctuations
(
Wiley
,
New York, NY
,
1977
).
30.
J.
Wang
,
C.
Li
, and
E.
Wang
, “
Potential and flux landscapes quantify the stability and robustness of budding yeast cell cycle network
,”
Proc. Natl. Acad. Sci. U. S. A.
107
,
8195
8200
(
2010
).
31.
J.
Wang
,
K.
Zhang
,
L.
Xu
, and
E.
Wang
, “
Quantifying the Waddington landscape and biological paths for development and differentiation
,”
Proc. Natl. Acad. Sci. U. S. A.
108
,
8257
8262
(
2011
).
32.
H.
Yan
et al, “
Nonequilibrium landscape theory of neural networks
,”
Proc. Natl. Acad. Sci. U. S. A.
110
,
E4185
E4194
(
2013
).
33.
X.
Fang
,
K.
Kruse
,
T.
Lu
, and
J.
Wang
, “
Nonequilibrium physics in biology
,”
Rev. Mod. Phys.
91
,
045004
(
2019
).
34.
J.
Wang
, “
Landscape and flux theory of non-equilibrium dynamical systems with application to biology
,”
Adv. Phys.
64
,
1
137
(
2015
).
35.
H.
Feng
,
B.
Han
, and
J.
Wang
, “
Adiabatic and non-adiabatic non-equilibrium stochastic dynamics of single regulating genes
,”
J. Phys. Chem. B
115
,
1254
1261
(
2011
).
36.
K.
Zhang
,
M.
Sasai
, and
J.
Wang
, “
Eddy current and coupled landscapes for nonadiabatic and nonequilibrium complex system dynamics
,”
Proc. Natl. Acad. Sci. U. S. A.
110
,
14930
14935
(
2013
).
37.
H.
Wang
,
X.
Li
, and
J.
Wang
, “
Quantifying the potential and flux landscapes for nonequilibrium multiverse, a new scenario for time arrow
,”
J. High Energy Phys.
2021
,
105
.
38.
T. L.
Hill
, “
Survey of the elements of free energy transduction
,” in
Free Energy Transduction and Biochemical Cycle Kinetics
(
Springer
,
1989
), pp.
1
37
.
39.
J.
Schnakenberg
, “
Network theory of microscopic and macroscopic behavior of master equation systems
,”
Rev. Mod. Phys.
48
,
571
(
1976
).
40.
D.
Jiang
,
M.
Qian
, and
M.
Qian
,
Mathematical Theory of Nonequilibrium Steady-States (Lecture Notes in Mathematics)
(
Springer
,
Berlin
,
2004
).
41.
X.
Luo
,
L.
Xu
,
B.
Han
, and
J.
Wang
, “
Funneled potential and flux landscapes dictate the stabilities of both the states and the flow: Fission yeast cell cycle
,”
PLoS Comput. Biol.
13
,
e1005710
(
2017
).
42.
B.
Han
and
J.
Wang
, “
Least dissipation cost as a design principle for robustness and function of cellular networks
,”
Phys. Rev. E
77
,
031922
(
2008
).
43.
J.
Wang
,
L.
Xu
,
E.
Wang
, and
S.
Huang
, “
The potential landscape of genetic circuits imposes the arrow of time in stem cell differentiation
,”
Biophys. J.
99
,
29
39
(
2010
).
44.
C.
Li
,
E.
Wang
, and
J.
Wang
, “
Landscape, flux, correlation, resonance, coherence, stability, and key network wirings of stochastic circadian oscillation
,”
Biophys. J.
101
,
1335
1344
(
2011
).
45.
A.
Nern
and
R. A.
Arkowitz
, “
G proteins mediate changes in cell shape by stabilizing the axis of polarity
,”
Mol. Cell
5
,
853
864
(
2000
).
46.
C. W.
Gardiner
,
Handbook of Stochastic Methods
(
Springer
,
Berlin
,
1985
), vol.
3
.
47.
N.
Van Kampen
,
Stochastic Processes in Physics and Chemistry
(
Elsevier
,
Amsterdam
,
2007
).
48.
L.
Xu
,
H.
Shi
,
H.
Feng
, and
J.
Wang
, “
The energy pump and the origin of the non-equilibrium flux of the dynamical systems and the networks
,”
J. Chem. Phys.
136
,
165102
(
2012
).
49.
B.
Han
and
J.
Wang
, “
Quantifying robustness and dissipation cost of yeast cell cycle network: The funneled energy landscape perspectives
,”
Biophys. J.
92
,
3755
3763
(
2007
).
50.
W.
Krauth
,
Statistical Mechanics: Algorithms and Computations
(
OUP Oxford
,
2006
), Vol.
13
.
51.
Y.
Cao
,
H.-M.
Lu
, and
J.
Liang
, “
Probability landscape of heritable and robust epigenetic state of lysogeny in phage lambda
,”
Proc. Natl. Acad. Sci. U. S. A.
107
,
18445
18450
(
2010
).
52.
M.
Qian
and
Z.
Hou
,
Reversible Markov Process
(
Hunan Scientific and Technical Press
,
Changsha
,
1979
).
53.
R. K. P.
Zia
and
B.
Schmittmann
, “
Probability currents as principal characteristics in the statistical mechanics of non-equilibrium steady states
,”
J. Stat. Mech.
2007
,
P07012
.
54.
Z. D.
Zhang
and
J.
Wang
, “
Curl flux, coherence, and population landscape of molecular systems: Nonequilibrium quantum steady state, energy (charge) transport, and thermodynamics
,”
J. Chem. Phys.
140
,
245101
(
2014
).
55.
H.
Qian
,
Open-System Nonequilibrium Steady State: Statistical Thermodynamics, Fluctuations, and Chemical Oscillations
(
ACS Publications
,
2006
), pp
15063
15074
.
56.
K.-Y.
Kim
and
J.
Wang
, “
Potential energy landscape and robustness of a gene regulatory network: Toggle switch
,”
PLoS Comput. Biol.
3
,
e60
(
2007
).
57.
D. T.
Gillespie
, “
Stochastic simulation of chemical kinetics
,”
Annu. Rev. Phys. Chem.
58
,
35
55
(
2007
).
58.
H.
Ge
and
H.
Qian
, “
Physical origins of entropy production, free energy dissipation, and their mathematical representations
,”
Phys. Rev. E
81
,
051133
(
2010
).
59.
H.
Feng
and
J.
Wang
, “
Potential and flux decomposition for dynamical systems and non-equilibrium thermodynamics: Curvature, gauge field, and generalized fluctuation-dissipation theorem
,”
J. Chem. Phys.
135
,
234511
(
2011
).
60.
G. E.
Crooks
, “
Nonequilibrium measurements of free energy differences for microscopically reversible Markovian systems
,”
J. Stat. Phys.
90
,
1481
1487
(
1998
).
61.
G. E.
Crooks
, “
Entropy production fluctuation theorem and the nonequilibrium work relation for free energy differences
,”
Phys. Rev. E
60
,
2721
(
1999
).
62.
J.
Wang
,
K.
Zhang
, and
E.
Wang
, “
Kinetic paths, time scale, and underlying landscapes: A path integral framework to study global natures of nonequilibrium systems and networks
,”
J. Chem. Phys.
133
,
125103
(
2010
).
63.
F.
Zhang
,
L.
Xu
,
K.
Zhang
,
E.
Wang
, and
J.
Wang
, “
The potential and flux landscape theory of evolution
,”
J. Chem. Phys.
137
,
065102
(
2012
).
64.
J.
Wang
,
B.
Huang
,
X. F.
Xia
, and
Z. R.
Sun
, “
Funneled landscape leads to robustness of cell networks: Yeast cell cycle
,”
PLoS Comput. Biol.
2
,
1385
1394
(
2006
).
65.
L.
Xu
and
J.
Wang
, “
Curl flux as a dynamical origin of the bifurcations/phase transitions of nonequilibrium systems: Cell fate decision making
,”
J. Phys. Chem. B
124
,
2549
2559
(
2020
).
66.
S.
Ghaemmaghami
et al, “
Global analysis of protein expression in yeast
,”
Nature
425
,
737
741
(
2003
).
67.
W.-K.
Huh
et al, “
Global analysis of protein localization in budding yeast
,”
Nature
425
,
686
(
2003
).
68.
R.
Howson
et al, “
Construction, verification and experimental use of two epitope-tagged collections of budding yeast strains
,”
Int. J. Genomics
6
,
474836
(
2005
).
69.
T.
Kruitwagen
,
P.
Chymkowitch
,
A.
Denoth-Lippuner
,
J.
Enserink
, and
Y.
Barral
, “
Centromeres license the mitotic condensation of yeast chromosome arms
,”
Cell
175
,
780
(
2018
).
70.
F.
Zucca
,
C.
Visintin
,
J.
Li
,
S. P.
Gygi
, and
R.
Visintin
, “
APC/CCdc20-mediated degradation of Clb4 prompts astral microtubule stabilization at anaphase onset
,”
J. Cell Biol.
222
,
e202203089
(
2023
).
71.
A.
Ruggiero
,
Y.
Katou
,
K.
Shirahige
,
M.
Séveno
, and
S.
Piatti
, “
The phosphatase PP1 promotes mitotic slippage through Mad3 dephosphorylation
,”
Curr. Biol.
30
,
335
(
2020
).

Supplementary Material