Abstract
Conduction delay and failure behaviors of action potentials with a high frequency along nerve fiber are related to the abnormal functions. For instance, upregulation of a hyperpolarization-activated cation current (Ih) is identified to reduce the conduction delay to recover the temporal encoding, and downregulation of the Ih current to enhance the conduction failure rate to ease the pain sensation, with the dynamic mechanisms remaining unclear. In the present paper, the dynamic mechanism is obtained in a chain network model with coupling strength (gc) and action potentials induced by periodic stimulations with a period (Ts). At first, as the action potentials exhibit a high frequency corresponding to a short Ts and the network has a small gc, i.e., a short and unrecovered afterpotential and a small coupling current, the conduction delay is reproduced. The conduction failure is reproduced for Ts shorter and gc smaller than those of the conduction delay, presenting a direct relationship between the two behaviors. Then, the conduction delay and failure are explained with the response time and current threshold of an action potential evoked from the unrecovered afterpotential. The prolonged response time for short Ts and small gc presents the cause for the conduction delay, and the enhanced threshold for shorter Ts and smaller gc presents the cause for the conduction failure. Furthermore, reduction of the delay and enhancement of the failure rate respectively induced by upregulation and downregulation of the Ih current are reproduced and explained. The positive Ih current induces Hopf bifurcation advanced and resting membrane potential elevated. Then, upregulation and downregulation of the Ih current induce the afterpotential elevated to shorten the response time and reduced to enhance the threshold, respectively. The results present nonlinear dynamics for the non-faithful conduction behaviors and dynamical mechanism for the modulation effect of the Ih current on the conduction delay and failure related to encoding and pain.
Keywords: Spatiotemporal dynamics, Neuronal network, Propagation of action potential, Hyperpolarization-activated cation current (Ih), Bifurcation, Pain
Introduction
It is well-known that the generation of an action potential of a neuron is related to the threshold, which is associated with the complex nonlinear dynamics such as bifurcations (Duan et al. 2020; Guo et al. 2021; Xing et al. 2023; Lyu et al. 2023). The conduction of action potentials along an axon or nerve fiber exhibits complex spatiotemporal dynamics (Zhang et al. 2020; Ding and Jia 2021; Zang et al. 2022), playing important roles in information transmission (Freeman 2000; Shimba et al. 2021; Yao et al. 2022). In general, action potentials faithfully propagate along an axon, which corresponds to normal functions of the nervous system (Bucher and Goaillard 2011). However, complex conduction behaviors of action potentials, such as conduction delay or conduction failure behaviors, may correspond to abnormal functions or diseases. For example, the conduction delay behavior means that the conduction time of action potentials increases with respect to time, i.e., delayed action potentials appear, which are associated with the changes in temporal encoding and dysfunction in learning, thinking, and acting (Ballo et al. 2012; Arancibia-Cárcamo et al. 2017; Zhang et al. 2017). Some action potentials fail to propagate along an axon, forming the conduction failure behavior, which is associated with pathological pain, demyelinating disease, and so on (Lesnick et al. 2007; Wang et al. 2016a; Waxman 2006; Wimmer et al. 2010). Then, identification of mechanism for the spatiotemporal dynamics of different conduction behaviors is helpful to modulate the related abnormal functions.
Conduction delay means that the conduction velocity of action potentials decreases with respect to time (Bucher and Goaillard 2011), which is related to the frequency of action potentials. In some neurological disorders such as the demyelinating disease, conduction velocity of action potentials significantly reduces (Waxman 1977; Kiziltan et al. 2007; Dutta et al. 2018), resulting in altered temporal encoding of information contained in the action potentials (Panzeri et al. 2010). It is well-known that the type or diameter of the nerve fibers determines the conduction velocity. Conduction delay behavior appears as a burst containing multiple action potentials with high frequency propagates along the pyloric dilator axon of lobsters, which is related to the digestive function (Zhang et al. 2017). To investigate the conduction delay behavior, current pulse stimulation with high frequency is used to evoke action potentials. The conduction time (delay) becomes long with increasing the frequency of action potentials (Ballo et al. 2012). A special ionic current, hyperpolarization-activated cation current (called Ih current for convenience)) mediated by the hyperpolarization-activated cyclic-nucleotide-gated (HCN) channels (Lezmy et al. 2021; Khurana et al. 2011; Byczkowicz et al. 2019), plays important roles in influencing the conduction time. Application of dopamine to enhance the Ih current can reduce the conduction time, resulting in enhanced temporal fidelity of action potentials, which may be beneficial for the temporal encoding (Ballo and Bucher 2009; Ballo et al. 2012; Zhang et al. 2017). In many studies on the unmyelinated axon and cerebellar mossy fiber, upregulation of the Ih current can increase the conduction velocity of action potentials, i.e., reduces the conduction delay, and blockage of the Ih current significantly enhances the conduction time of action potentials (Byczkowicz et al. 2019; Roth and Hu 2020), i.e., enhances the conduction delay. These experimental observations show that upregulation of the Ih current is an important modulation target to reduce the conduction delay to recover the normal function. Unfortunately, it is unclear how the Ih current affects the conduction delay behavior. Then, identification of dynamics of the conduction delay modulated by the frequency of action potentials, conduction velocity of the axon, and the Ih current is one objective of the present paper.
Another non-faithful conduction behavior, conduction failure behavior, has been observed in more and more experiments (Zhu et al. 2009; Dilley et al. 2013; Wang et al. 2016b; Hamada et al. 2017; Mao et al. 2019). There are various factors to influence the conduction failure behavior (Bucher and Goaillard 2011; Ratas and Pyragas 2012; Wang et al. 2016b), such as the type of fibers to determine the conduction velocity, the frequency of action potentials, ionic currents, and so on. The conduction velocity of unmyelinated axons is often slower than that of myelinated axons (Gillespie and Stein 1983). For unmyelinated axons, the larger the diameter is, the faster the conduction velocity is, due to that the conduction velocity is positively correlated with the diameter of axon (Hodgkin 1954). The conduction failure behavior is often observed in the unmyelinated axon such as the C-fiber (Sun et al. 2012; Ballo et al. 2012; Zhang et al. 2017). For example, a representative of the conduction failure behavior is associated with diabetes-induced pain information, which is reduced as the conduction failure is enhanced (Sun et al. 2012). The result suggests that the conduction failure behavior may be an intrinsic self-inhibitory mechanism of the pain sensation. Conduction failure behavior has also been confirmed in the C-fibers of rabbits (Zhu et al. 2009), especially for the action potentials with a high frequency induced by an external stimulus with a short period (Ratas and Pyragas 2012; Cross and Robertson 2016). Ion currents such as the Ih, potassium (K+), and pump currents play important roles in regulating the conduction failure behavior (Crotty and Levy 2007; Kueh et al. 2016; Wang et al. 2016a; Wang et al. 2016b). Especially, the Ih current plays important roles in inflammatory and neuropathic pain (Chaplan et al. 2003; Weng et al. 2012; Jansen et al. 2021; Bernard Healey et al. 2021). For instance, blockage of the Ih current significantly increases the conduction failure (Zhu et al. 2009), which has an analgesic effect (suppressing the pain sensation) (Chaplan et al. 2003; Wang et al. 2016a). Then, identifying the dynamics of conduction failure modulated by the frequency of action potentials, the conduction velocity, and the Ih current is the second objective of the present paper. In addition, as the conduction time of action potentials increases to a certain extent (Zhu et al. 2009), the following action potential fails to conduct, forming the conduction failure behavior, which presents a possible relationship between the conduction delay and failure behaviors. Then, identification of the direct relationship between the conduction delay and failure behaviors is another objective of the present paper.
The theoretical models of reaction-diffusion equations or chain network are used to simulate the axon (Ding and Jia 2021; Zhang et al. 2019b). The complex spatiotemporal behaviors related to the conduction behaviors are reproduced in various studies, and the spatiotemporal behaviors change with respect to multiple modulation parameters (Zhang et al. 2019c, 2020; Song et al. 2019; Zang and Marder 2021; Zang et al. 2022). For example, the influences of the nerve fiber type, ionic currents, temperature, and memristor on the conduction behaviors are investigated. For the chain network model containing multiple compartments to simulate the axon, a compartment receives coupling pulse current from the neighborhood compartments at a phase of the afterpotential (membrane potentials following an action potential). Then, three factors, the amplitude of the coupling current pulse, the timing or application phase of the pulse, and the level of the afterpotential at the application phase determine the conduction delay or failure behaviors. The application phase is determined by the frequency of action potentials (Zhang and Gu 2019), the amplitude is determined by the coupling strength, and the level of afterpotential is determined by the factors of the compartment such as the Ih current. For instance, conduction delay behavior appears if a delayed action potential is induced by the coupling current from the afterpotential, and conduction failure occurs if an action potential cannot be caused sometimes. Then, with a single neuron model, the threshold of current to evoke an action potential is used to characterize the conduction failure behavior (Zhang et al. 2020), which has not been used to study the modulation effect of the Ih current on the conduction failure. Unfortunately, the dynamical mechanism for the conduction delay behavior remains unclear. To present an effective measure to characterize the dynamics underlying the conduction delay behavior is one goal of the present paper. Then, identification of dynamical mechanisms for the conduction delay with a novel measure and the conduction failure with the current threshold is the last objective of the present paper. Especially, the modulation effects of the Ih current on the conduction delay and failure behaviors are studied.
In this article, the experimental results of the conduction delay and conduction failure behaviors are reproduced in a chain network model with the Ih current (Hodgkin and Huxley 1952; Khurana et al. 2011), and the dynamical mechanisms are presented with nonlinear responses of the afterpotentials in a neuron model, which is helpful to modulate the temporal encoding and pathological pain. Firstly, conduction delay behavior similar to the experimental observations is reproduced in a chain network with the coupling strength labeled as gc. External pulse stimulations with a period (labeled as Ts) are used to evoke action potentials at one end of the chain. For a short Ts to evoke action potentials with a high frequency, conduction delay behavior appears for a relatively small gc, due to the unrecovered afterpotential within the short Ts. Conduction time to reflect the conduction delay becomes short as the Ih current increases. When gc is large or Ts is long, faithful conduction behavior occurs. Secondly, conduction failure behavior resembling experimental observations (Wang et al. 2016a) is reproduced, appearing for a shorter Ts and a lower gc, compared with the conduction delay behavior, which presents a direct relationship between conduction delay and failure behaviors. As the Ih current decreases, the conduction failure rate increases in wide ranges of Ts and gc, which contains four cases. Finally, the mechanisms for the conduction delay and failure behaviors are acquired in a single neuron model. The conduction delay and failure are respectively explained with the response time (a novel indicator used in the present paper) and current threshold for an action potential, which is caused by a current pulse applied to different phases of the unrecovered afterpotentials. A small gc induces a small coupling current pulse which cannot go beyond the threshold sometimes to evoke action potential, presenting the cause for the conduction failure appearing at small gc. The response time becomes long and the current threshold becomes large with decreasing the application phase of the stimulation current pulse, which can explain the conduction delay and failure appearing at a short Ts. Especially, the Ih current induces Hopf bifurcation advanced and afterpotential related to the Hopf bifurcation elevated, resulting in a shortened response time and reduced current threshold, which can explain the modulation effects of Ih current. The results present dynamical mechanisms for the conduction delay and conduction failure behaviors modulated by the Ih current, which is helpful to understand the viewpoint that the Ih current may be used as a target to regulate the conduction delay and conduction failure, such as upregulation of the Ih current to reduce conduction delay to improve the temporal encoding and blockage of the Ih current to enhance the conduction failure rate to ease pain.
The rest of this article is organized as follows. Sects. ”Model and method”, “Result”, and “Conclusion and Discussion” present the Model and Method, Result, and Conclusion, respectively.
Model and method
Chain network model to simulate axon
The axon is described by a chain network model containing multiple compartments and the compartment in the network is described by Hodgkin-Huxley model (HH) with the Ih current (Hodgkin and Huxley 1952; Khurana et al. 2011). The network model is described as follows:
| 1 |
| 2 |
| 3 |
| 4 |
| 5 |
where 100 compartments are considered in the network, and “i” for a compartment denotes its sequential number in the chain network. Each compartment except the first and 100-th ones is connected to the two neighboring compartments, the first compartment is only coupled to the second one, and the compartment 100 to the compartment 99. Then, the coupling current in Eq. (1) is Ic,i = gc (Vi-1 + Vi+1 − 2Vi) (1 < i < 100), Ic,1 = gc (V2 − V1), and Ic,100 = gc (V99 − V100), where gc denotes the coupling strength. If Ic,i in Eq. (1) and the sequential numbers “i” in Eqs. (1–5) are ignored, Eqs. (1–5) become the HH model to describe a single neuron, which is introduced as follows:
V, m, h, and n represent the membrane potential, activation, and inactivation of sodium (Na+) channel, and activation of potassium (K+) channel, respectively. The parameter C is capacitance of membrane, and Iapp is the applied current. The currents INa, IK, IL, and Ih represent the Na+ current, K+ current, leak current, and hyperpolarization-activated cation current, respectively, which are given as follows:, , , and . The parameters gNa, gK, gL, and gh are the maximal conductances, and ENa, EK, EL, and Eh denote the equilibrium potentials. The related functions are given by: , , , , , , , and .
Ih current
Especially, for a single neuron or compartment, the Ih current is considered in the present paper, where H, gh, and Eh represent the open probability of the inactivation gate, the maximum conductance, and the equilibrium potential of the Ih current, respectively.
External voltage pulse stimulus
Axon at resting state is chosen as control in the biological experiments, then, one end of the axon (stimulation region) is stimulated by external periodic voltage stimulation (labeled with Sti) to evoke action potentials conducting to the other end (conduction region) (Zhu et al. 2009), as depicted in Fig. 1. The voltage pulse stimulation (please refer to Figs. 7a, 13) similarly to the experiments (Zhu et al. 2009) is applied in the present paper. Pulse width 2 ms and pulse strength 15 mV is used, and then a pulse can induce an action potential. The period of the voltage pulses is labeled as Ts and is set as a control parameter to determine the frequency of action potentials. Compartments 1 to 5 are the stimulation region in the present paper, and the remaining compartments are the conduction region.
Fig. 1.

Sketch map of action potentials that are evoked at one end (stimulation region) by external voltage pulses and conduct along the axon (conduction region)
Fig. 7.
Conduction behaviors. a Conduction failure at gc = 0.08 mS/cm2; b enlargement of the first six spikes of the panel (a); c conduction time for conduction failure behavior at gc = 0.08 (black) and 0.087 (red) mS/cm2; d conduction block behavior at gc = 0.05 mS/cm2
Fig. 13.
Small (gc = 0.06 mS/cm2, left column) and large (gc = 0.1 mS/cm2, right column) coupling currents respectively for the conduction failure and conduction delay. a1 and b1: External periodic pulse stimulation; a2 and b2: V5; a3 and b3: Coupling current Ic,6; a4 and b4: V6. Other parameter values: a short Ts (= 17 ms) and gh = 0.3 mS/cm2
Parameter values and method
In the present paper, C = 1 μF/cm2, gNa, gK, and gL are 120, 36, and 0.3 mS/cm2, respectively, ENa, EK, EL, and Eh are 50, − 77, − 59, and − 30 mV, respectively. Iapp, gh, Ts, and gc are the control parameters.
Fourth Runge–Kutta method is used to integrate the equations, and the integration time step is 0.001 ms. Software XPPAUT is employed to calculate the bifurcations (Cartwright 2010).
Result
Dynamics of conduction delay behavior
The period of external pulse stimulation to evoke action potentials in the stimulation region is labeled as Ts, the coupling strength between compartments of the chain network is labeled as gc, and the conductance of the Ih current is labeled as gh.
A representative of conduction delay behavior
For the conduction delay behavior, the conduction time of action potentials increases with respect to time. A representative of the conduction delay behavior for gc = 0.12 mS/cm2, Ts = 18 ms, and gh = 0 mS/cm2 is depicted in Fig. 2a, b. The circular column (left) represents the axon, where the blue region indicates the stimulation region and the gray area denotes the conduction region, and changes of three action potentials for V5 and V95 are shown as representatives here. For the j-th action potential (j is the sequential number of the action potentials), the conduction time from the compartment 5 to the compartment 95 is labeled as Dj. For Fig. 2a with gh = 0 mS/cm2, D1 = 164.14 ms, D2 = 173.79 ms, and D3 = 180.89 ms, showing that the conduction delay behavior appears. In the network model, Iapp = −4 μA/cm2 is used.
Fig. 2.
Conduction delay behaviors. a gh = 0 mS/cm2; b gh = 0.3 mS/cm2; c Comparison between panels (a) and (b); d Enlargement of V95 in panels (a) and (b)
The conduction delay at gh = 0.3 mS/cm2 is illustrated in Fig. 2b. D1 = 153.60 ms, D2 = 157.92 ms, and D3 = 161.49 ms, which are shorter than those of gh = 0 mS/cm2. The comparison between panels (a) and (b) is shown in Fig. 2c. For the 5-th compartment, action potential for gh = 0 mS/cm2 appears at the same time as that of 0.3 mS/cm2. However, for the 95-th compartment, the action potential for gh = 0.3 mS/cm2 (blue) appears earlier, indicating a faster conduction velocity, compared with gh = 0 mS/cm2 (black). As illustrated in Fig. 2d, the local enlargement of V95 shows that the resting membrane potential (before the first action potential) for gh = 0.3 mS/cm2 is higher, compared with 0 mS/cm2. For both gh values, the afterpotential does not recover to the resting potential, due to that Ts = 18 ms is too short.
The influence of the Ih current on conduction delay behavior
For gh = 0.3 mS/cm2, Dj labeled with blue circles gradually increases with respect to time and tends to be a steady value 176.75 ms (after t ≈ 360 ms), as shown in Fig. 3a (gc = 0.12 mS/cm2 and Ts = 18 ms). The time interval between circles in Fig. 3a is Ts, and “t” of the x-axis for the j-th circle is j × Ts, which corresponds to the application time of the stimulation pulse to evoke the j-th action potential of the compartment 5. For gh = 0 mS/cm2, Dj labeled with black squares gradually increases and tends to be a steady value of 223.01 ms (after t ≈ 594 ms). The conduction time for gh = 0.3 mS/cm2 becomes shorter, compared with 0 mS/cm2, indicating that increasing Ih current can reduce the conduction time, which is consistent with the experimental results (Ballo et al. 2012). As can be found from Fig. 3a, the conduction time converges to a steady value, showing that the convergence process of the conduction time corresponds to a transient behavior. The convergence processes and the steady values of the convergence time for different values of gh are different.
Fig. 3.
Changes of conduction time for gh = 0 (black) and 0.3 mS/cm2 (blue). a Dj, b ΔDj
For gh = 0 mS/cm2, ΔDj = Dj+1 − Dj (j = 1, 2, 3, …) is shown by a black curve in Fig. 3b, which reflects the speed of change in the conduction time with respect to the application time of the stimulation. ΔDj decreases from a positive value to zero gradually, showing that the speed of change in the conduction time becomes small. ΔDj for gh = 0.3 mS/cm2 (blue) is smaller, compared with 0 mS/cm2 (black), as illustrated in Fig. 3b. The results show that increases of the Ih current induces the speed of change in the conduction time decreased.
Conduction delay for a short Ts and faithful conduction for a long Ts
When gh = 0.3 and gc = 0.1 mS/cm2, conduction delay behavior occurs at a shorter Ts (18 ms), as shown in Fig. 4a. Two spikes are chosen as a representative, and the conduction time is 153.6 ms and 157.9 ms, respectively. However, for a large Ts (50 ms), faithful conduction behavior occurs, as depicted in Fig. 4b. The two conduction times are both 153.6 ms, showing that the conduction time does not change with increasing time. The unchanged conduction time for Ts = 50 ms is depicted by the bottom curve (dark blue) in Fig. 4c. The convergence processes of the conduction time for a short Ts are shown by the top four curves in Fig. 4c (Ts = 18, 19, 20, and 21 ms from top to bottom). For the conduction delay behavior at a small Ts, conduction time increases at first, and then tends to be a steady value. The conduction time is long for a small Ts.
Fig. 4.
Conduction behaviors at different Ts values. a Conduction delay for Ts = 18 ms; b faithful conduction for Ts = 50 ms; c changes of conduction time at different values of Ts; d V5 for Ts = 18 ms in panel (a) (Black solid curve) and for Ts = 50 ms in panel (b) (Red dotted curve); e Enlargement of panel (d)
The local enlargement of V5 for Ts = 18 ms shown in panel (a) and for Ts = 50 ms shown in panel (b) are depicted by black solid and red dotted curves in Fig. 4d, respectively. The first spikes for Ts = 18 ms and Ts = 50 ms are duplicated, because the first stimulation times are the same (at 11 ms). The local enlargement of Fig. 4d is shown in Fig. 4e. For Ts = 18 ms, the afterpotential (solid black curve) before the second stimulation (at 29 ms), as denoted by the former arrow (black), does not recover. For Ts = 50 ms, the afterpotential before the latter arrow (red) as well as a red dot (at ~ 53 ms) does not recover, while recovers after 53 ms. The second stimulation is applied at 61 ms, then, the membrane potential at the second stimulation recovers. The different levels of the afterpotentials at the second stimulation times are the cause for the conduction delay behavior or the faithful conduction behavior, which is addressed in SubSect. ”Response time curve to explain the conduction delay behavior” and Fig. 15.
Fig. 15.
The response time of an action potential evoked from the unrecovered afterpotentials to a current pulse stimulus applied at a phase Δt (pulse width of 2 ms). At different gh values (the black and blue curves respectively represent gh = 0 and 0.3 mS/cm2): a1 the response for Δt = 15 ms with pulse intensity ΔA of 8.5 μA/cm2; a2 changes of the response time. At different ΔA values (red and blue for ΔA = 8.5 and 9.5 μA/cm2, respectively): b1 the response for Δt = 15 ms and gh = 0 mS/cm2; b2 changes of the response time
Different conduction behaviors for different gc values
At different values of gc, changes of the conduction time exhibit different cases, showing different conduction behaviors, as illustrated in Fig. 5 (Ts = 18 ms and gh = 0.3 mS/cm2).
Fig. 5.

Faithful conduction behavior for a large value of gc (gc = 3 mS/cm2, magenta), and conduction delay for relatively small gc values (black, red, green, blue, and cyan respectively represent gc = 0.1, 0.12, 0.14, 0.2, and 0.5 mS/cm2)
For a large gc, for instance, gc = 3 mS/cm2, the conduction time does not change with respect to time, as shown by magenta (bottom curve) in Fig. 5, corresponding to the faithful conduction behavior.
For relatively small values of gc, such as gc = 0.1, 0.12, 0.14, 0.2, and 0.5 mS/cm2, the convergence processes of the conduction time are depicted in Fig. 5. The conduction time increases at first and then tends to be a steady value, i.e., conduction delay behavior appears, as shown by the top 5 curves in Fig. 5. The larger the gc is, the shorter the conduction time is, and the shorter the duration of conduction delay behavior is.
In addition, the conduction failure behavior and conduction block behavior occur for small values of gc (such as 0.08 and 0.087 mS/cm2) and a much smaller gc (0.05 mS/cm2), respectively, which are addressed in detail in the following Subsection.
Large gh induces large Ih currents to reduce the conduction delay
As shown in Fig. 3a, the conduction time for gh = 0 mS/cm2 is longer than that of gh = 0.3 mS/cm2. The role of the Ih current in modulating the conduction delay is shown in Fig. 6, left column for gh = 0 mS/cm2 and right column for gh = 0.3 mS/cm2 (with Ts = 18 ms and gc = 0.12 mS/cm2). The voltage stimulation pulses applied to the compartment 5, V5, V6, Ih,6 (the Ih current of the compartment 6), Itotal,6 (the total current of the compartment 6), and enlargement of the Ih,6 current and the Itotal,6 current are shown from top to bottom rows, respectively. In Fig. 6a6, gh = 0 mS/cm2, the Ih,6 current (red) is 0 μA/cm2, and the Itotal,6 current (blue) for the afterpotential is relatively small (the maximum value is about 0.24 μA/cm2). In Fig. 6b6, gh = 0.3 mS/cm2, the maximum value of the Ih,6 current for the afterpotential is about 1.82 μA/cm2, and the Itotal,6 current is relatively large (the maximum value is about 0.52 μA/cm2). The result shows that the positive Ih current for the afterpotential plays a role in raising the total current, which can induce earlier appearance of the action potentials. Then, the conduction time of the first spike from the fifth compartment to the sixth compartment for gh = 0.3 mS/cm2 (about 1.96 ms) is shorter than that for gh = 0 mS/cm2 (about 2.43 ms). This result can explain that the conduction time for gh = 0.3 mS/cm2 is shorter than that for gh = 0 mS/cm2 in Fig. 3a.
Fig. 6.
The influence of the Ih currents on conduction delay for gh = 0 mS/cm2 (left column) and 0.3 mS/cm2 (right column). a1 and b1: External periodic pulse stimulations; a2 and b2: V5; a3 and b3: V6; a4 and b4: Ih,6 current; a5 and b5: Itotal,6 current; a6 and b6: enlargement of the Ih,6 current (red) and the Itotal,6 current (blue). Ts = 18 ms and gc = 0.12 mS/cm2
Dynamics of conduction failure behavior
Conduction failure behavior and conduction block behavior
A representative of conduction failure behavior is shown in Fig. 7a (Ts = 18 ms, gc = 0.08 mS/cm2, and gh = 0.3 mS/cm2). It can be found that each stimulation pulse (black labeled with “Sti”) in stimulation region can evoke an action potential for the compartment 5, V5. Here, 22 spikes are depicted as representatives due to the limited range of the time axe. For the first spike of the compartment 5, the conduction to the compartment 50 and to the compartment 95 is represented by the left line with an arrow in Fig. 7a. For the tenth spike of the compartment 5, the conduction to the compartment 50 and to the compartment 95 is shown by the middle line with an arrow. For the sixteenth spike of the compartment 5, the conduction to the compartment 50 is shown by the right line with an arrow, the conduction to the compartment 95 is not shown due to the limited range of the time axe. For the seventeenth to twenty-second spikes of the compartment 5, the conductions are not illustrated, due to the limited ranges of the time axe in Fig. 7a. As can be found in Fig. 7a, the 3 k-th (k = 1, 2, 3) action potentials (one out of three) of compartment 5 cannot conduct to the conduction region, resulting in a long interspike interval after the 2 k-th (k = 1, 2, 3) action potential of the compartment 50 and 95. To clearly show the dynamics of the conduction failure, the first six spikes of the compartment 5 are plotted in Fig. 7b as representatives. Correspondingly, 4 spikes conduct to the compartment 50 and the compartment 95 and 2 spikes fail to conduct. From the 5-th compartment to the 95-th compartment, the conduction time for the 1-st, 2-nd, 4-th, and 5-th spikes is 214.12 ms, 218.21 ms, 215.45 ms, and 219.19 ms, respectively. The conduction time for the second spike is larger than the first one, showing that the conduction velocity becomes slow. However, no conduction time can be acquired for the 3-rd spike (6-th spike), as depicted by the red dashed line, due to the conduction failure of the 3-rd spike (6-th spike) of the compartment 5.
The conduction time for the conduction failure behavior at gc = 0.08 (black) and 0.087 (red) mS/cm2 is illustrated in Fig. 7c. For gc = 0.08 mS/cm2 (black), there are gaps after the 2 k-th spikes (k = 1, 2, 3, …), due to the conduction failure shown in Fig. 7a, b. For gc = 0.087 mS/cm2, there are gaps after the 4 k-th spikes (k = 1, 2, 3, …), showing that the 5 k-th (k = 1, 2, 3, …) action potentials of the compartment 5 fail to conduct. Especially, before the conduction failure of a spike, the conduction time of action potentials gradually increases, similar to the experimental observations (Zhu et al. 2009). The smaller the gc is, the longer the conduction time is, and the higher the frequency of the action potentials failing to conduct is.
To distinguish the conduction failure for gc = 0.08 (black) and 0.087 (red) mS/cm2 in Fig. 7c, conduction failure rate for the compartment i denoted as Ri can be defined. As shown in Fig. 7b (gc = 0.08 mS/cm2), 4 out of 6 action potentials in the stimulation region conduct to the compartments 50 and 95, i.e., 2 out of 6 spikes fail to conduct. Then, the conduction failure rate is defined as R50 = (6–4)/6 = 1/3, and R95 = (6–4)/6 = 1/3. Similarly, R50 = 1/5 and R95 = 1/5 for gc = 0.08 mS/cm2 (black in Fig. 7c). More generally, if the number of the action potentials induced by the stimulation pulses applied in the stimulation region is labeled as M (M is an integer), and the number of the action potentials conducted to the compartment i is labeled as N (N is an integer between 0 and M), the conduction failure rate of the i-th compartment is defined as Ri = (M − N) / M, with a value between 0 and 1. 0 < Ri < 1 denotes that only some action potentials successfully conduct, i.e., conduction failure occurs. Ri = 0 means that no action potentials fail to propagate, which corresponds to conduction delay or faithful conduction behaviors. Ri = 1 means that all action potentials fail to conduct, called conduction block behavior, as shown in Fig. 7d, with gc further decreased to 0.05 mS/cm2.
The influence of Ts and gc on conduction behaviors
The compartment 95 is chosen as a representative to show the influence of gc and Ts on R95. For gh = 0 mS/cm2, the distribution of R95 on the parameter plane (Ts, gc) is depicted in Fig. 8a. Green, red, and blue represent 0 < R95 < 1 (conduction failure), R95 = 1 (conduction block), and R95 = 0 (conduction delay and faithful conduction for long Ts), respectively. The purple curve represents the border between the conduction failure and conduction delay (faithful conduction), and the magenta curve denotes the border between the conduction failure and block behaviors. Two borders are separated when Ts < ~ 37.5 ms and are near or coincide when Ts > ~ 37.5 ms. The conduction failure and delay behaviors appear for relatively small gc and short Ts, and the conduction block for very small gc. Especially, compared with the conduction delay behavior, the conduction failure behavior occurs for shorter Ts and smaller gc, which presents a direct relationship between the conduction delay and failure behaviors. A short Ts corresponds to a high frequency of the action potentials and a small gc to the C-fiber, showing that the conduction failure occurs for stimulations with a short Ts, resembling the experimental results of the C-fibers (Sun et al. 2012). With increasing gc, the range of Ts for the conduction failure decreases, i.e., presenting the changing regularity of the purple curve behavior. It is the faithful conduction behavior that locates up-right to the conduction delay behavior (not shown here).
Fig. 8.
The influence of Ts and gc on the conduction failure rate. a R95 for gh = 0 mS/cm2; b R95 for gh = 0.3 mS/cm2. Green, blue, and red represent 0 < R95 < 1, R95 = 0, and R95 = 1, respectively; c difference between panels (a) and (b), labeled as ∆R. ∆R = 0, 0 < ∆R < 1, and ∆R = 1 is shown by blue, light blue, and red, respectively
The influence of the Ih current on the conduction failure rate
For gh = 0.3 mS/cm2, the distribution of R95 is shown in Fig. 8b. Green, blue, and red areas respectively represent 0 < R95 < 1, R95 = 0, and R95 = 1. The black curve represents the border between 0 < R95 < 1 and R95 = 0, and the yellow curve denotes the border between R95 = 1 and 0 < R95 < 1 or R95 = 0. The conduction delay/faithful conduction region becomes larger for gh = 0.3 mS/cm2, compared with 0 mS/cm2, which indicates that increase of the Ih current facilitates conduction of the action potentials.
∆R, obtained by subtracting R95 for gh = 0.3 mS/cm2 from R95 for gh = 0 mS/cm2, can reflect the difference in failure rate. The distribution of ∆R on (Ts, gc) plane is depicted in Fig. 8c, which can reflect the changes of conduction failure rate induced by downregulation of the Ih current. ∆R = 0, 0 < ∆R < 1, and ∆R = 1 are shown by blue, light blue, and red, respectively. Combined with the borders between different conduction behaviors and borders between ∆R = 0, 0 < ∆R < 1, and ∆R = 1, (Ts, gc) plane is divided into 7 regions. ∆R = 0 appears in regions labeled with 5, 6, and 7 (not the focus of the present paper), 0 < ∆R < 1 in regions labeled with 1, 2, and 4, and ∆R = 1 in region 3. ∆R > 0 in regions 1, 2, 3, and 4 shows that the conduction failure rate enhances as gh decreases, labeled as case-1, -2, -3, and -4, respectively. The four cases show that the downregulation of gh enhances the conduction failure rate in wide ranges of (Ts, gc) plane, resembling the experiment results (Wang et al. 2016a). More details for the four cases are addressed in the following paragraphs.
Four cases of enhancement of the conduction failure rate
Case-1: enhanced failure rate for the conduction failure behavior
A representative for case-1 is shown in Fig. 9 (gc = 0.1 mS/cm2 and Ts = 17 ms). The spike trains of the compartment 6 for gh = 0.3 mS/cm2 exhibit R6 = 1/6, as shown in Fig. 9a1. As gh decreases to 0 mS/cm2, R6 increases to 1/2, as shown in Fig. 9b1. To present a global view, the spatiotemporal behaviors for gh = 0.3 and 0 mS/cm2 are depicted in Fig. 9a2, b2, respectively. The abscissa and ordinate denote the sequential number of compartments in the network and time, respectively. The left bright horizontal lines represent the peaks of action potentials in the stimulation regions (compartments 1 to 5), and the blue regions between two bright horizontal lines represent the membrane potential (afterpotential) between peaks of the action potentials. The bright oblique lines represent peaks of the action potentials in the conduction region (compartments 6 to 100), and the “oblique” reflects the propagations of the action potentials. Compared with stimulation region, the disappearance of bright lines in conduction region corresponds to conduction failure of action potentials. As depicted in Fig. 9a2, one out of six action potentials fail to conduct. As depicted in Fig. 9b2, one of very two action potentials fails to conduct.
Fig. 9.
Downregulation of gh causes enhancement of the failure rate. Left for gh = 0.3 mS/cm2: a1 spike trains with R6 = 1/6; a2 spatiotemporal behaviors. Right for gh = 0 mS/cm2: b1 spike trains with R6 = 1/2; b2 spatiotemporal behaviors
Case-2: from conduction delay to conduction failure
A representative of case-2 for gc = 0.12 mS/cm2 and Ts = 17 ms is depicted in Fig. 10. Conduction delay and conduction failure behaviors occur for gh = 0.3 and 0 mS/cm2, respectively, and the corresponding conduction failure rates are 0 and 0.2, respectively. Figure 10a1, b1 show the stable behavior of membrane potential of the 6-th compartment for gh = 0.3 and 0 mS/cm2, respectively, and Fig. 10a2 and Fig. 10b2 show the spatiotemporal behaviors of action potentials. The results show that the downregulation of gh can enhance the failure rate. Here, Fig. 10a1 shows the membrane potential of conduction delay behavior, when the conduction time tends to be a constant value.
Fig. 10.
Downregulation of gh induces conduction delay behavior changed to conduction failure. Left for gh = 0.3 mS/cm2: a1 spike trains with R6 = 0; a2 spatiotemporal behaviors. Right for gh = 0 mS/cm2: b1 spike trains with R6 = 1/5; b2 spatiotemporal behaviors
Case-3: from conduction delay to conduction block
When gc = 0.06 mS/cm2 and Ts = 30 ms, the conduction delay behavior (left column) changes to the conduction block behavior (right column) as gh decreases from 0.3 to 0 mS/cm2, as illustrated in Fig. 11. Membrane potentials of the compartment 6 are depicted in upper row and spatiotemporal behaviors in the lower row. Similar to Fig. 10a1, Fig. 11a1 shows the membrane potential of the conduction delay behavior, when the conduction time tends to be a constant value.
Fig. 11.
Downregulation of gh induces conduction delay changed to conduction block. Left for gh = 0.3 mS/cm2: a1 spike trains with R6 = 0; a2 spatiotemporal behaviors; Right for gh = 0 mS/cm2: b1 membrane potential for compartment 6; b2 spatiotemporal behaviors
Case-4: from conduction failure to conduction block
For gc = 0.06 mS/cm2 and Ts = 17 ms, as illustrated in Fig. 12, the conduction failure behavior (left column) changes to the conduction block behavior (right column) as gh changes from 0.3 to 0 mS/cm2. Membrane potentials of the compartment 6 are depicted in the upper row and spatiotemporal behaviors are in the lower row. As shown in Fig. 12a1, R6 = 1/2.
Fig. 12.
Downregulation of gh induces conduction failure changed to conduction block. Left for gh = 0.3 mS/cm2: a1 spike trains with R6 = 1/2; a2 spatiotemporal behaviors. Right for gh = 0 mS/cm2: b1 membrane potential for the compartment 6; b2 spatiotemporal behaviors
Small gc induces small coupling current to cause conduction failure from unrecovered afterpotentials for short Ts
As can be found in Fig. 8, conduction failure behavior occurs for smaller gc and shorter Ts, compared with the conduction delay behavior. For example, for a short Ts (17 ms), the conduction failure occurs at gc = 0.06 mS/cm2 and conduction delay (conduction time tends to be a constant value) occurs at gc = 0.1 mS/cm2 as illustrated in the left and right columns of Fig. 13, respectively. The voltage stimulation pulses applied to the compartment 5, V5, coupling current Ic,6, and V6 are shown from top to bottom rows, respectively. For the conduction failure behavior, the second, fourth, and sixth actional potentials fail to conduct, as shown by the stars in Fig. 13a4. The maximal value of the coupling current pulse is relatively small, about 6.0 μA/cm2, as illustrated in Fig. 13a3. However, all action potentials conduct for the conduction delay behavior, as illustrated by blue arrows in Fig. 13b4. The maximal value of the coupling current is relatively large, about 9.7 μA/cm2, as shown in Fig. 13b3. Then, it is the strength of the coupling current that determines the appearance of the conduction failure or conduction delay behaviors at a fixed Ts value. A large coupling current for a large gc can evoke action potentials to form the conduction delay behavior, while a small coupling current for a small gc sometimes cannot evoke action potentials sometimes to form the conduction failure behavior. Furthermore, it can be found that the afterpotential elevates when receiving the coupling pulse current, showing that the afterpotential has not recovered to the resting membrane potential within the short Ts. Then, responses of the unrecovered afterpotential to a current pulse stimulation with different strengths (corresponding to gc) and different application phases (corresponding to Ts), and the influence of gh on the responses, are helpful to explain the conduction failure behavior modulated by gc, Ts, and gh, which are addressed in SubSect. ”Mechanisms for modulations to the conduction delay and failure behaviors”.
Mechanisms for modulations to the conduction delay and failure behaviors
As mentioned above, both the conduction delay behavior and conduction failure behavior are dependent on strength of the coupling current pulse, phase of the unrecovered afterpotential when stimulated by the current pulse, and level of the unrecovered afterpotential. In fact, the level of the afterpotential is determined by gh, the stimulation time/phase of the current pulse is related to Ts, and the strength of the coupling current pulse is modulated by gc. Therefore, in the present Subsection, conduction delay and conduction failure behaviors in the network are explained with a single neuronal model with the Ih current, which is stimulated by a current pulse to simulate the coupling current. The strength ΔA and application time Δt of the pulse correspond to gc and Ts, respectively. Considering that the conduction delay behavior corresponds to delayed action potentials, response time of an action potential is proposed to characterize the conduction delay behavior. In addition, the current threshold to evoke an action potential is used to explain the conduction failure behavior. Then, two important curves to describe the responses of action potential and the modulations of ΔA, Δt, and gh on the two curves, are acquired. One curve to describe the response time of an action potential evoked from the afterpotential at different phases, which is called the response time curve, can well explain the conduction delay behavior. The other curve to describe the current strength threshold to induce an action potential from the afterpotential at different phases, i.e., the current threshold curve, can well explain the conduction failure behavior. Especially, the modulations of the Ih current on the afterpotential related to a Hopf bifurcation and on the two curves are obtained.
Ih current induces Hopf bifurcation advanced and afterpotential elevated
Here, the dynamics of HH model of a single neuron in the absence of external stimuli is studied. As depicted in Fig. 14a, black and blue curves represent the bifurcations near the steady state for gh = 0 and 0.3 mS/cm2, respectively. It is well-known that there is a subcritical Hopf bifurcation (H) and a saddle node bifurcation of limit cycles (SNLC). As gh increases from 0 to 0.3 mS/cm2, point H changes from Iapp ≈ 3.75 to 3.29 μA/cm2, and SNLC changes from Iapp ≈ 2.49 to 1.76 μA/cm2. With increasing gh, the Hopf bifurcation advances, due to that the Ih current is positive.
Fig. 14.
Dynamics of the HH model as gh changes from 0 (black) to 0.3 (blue) mS/cm2. a Bifurcations with respect to Iapp. Bold and dotted solid curves respectively for stable and unstable focus, and thin solid and dashed curves for stable and unstable limit cycles, respectively. b The resting potential for Iapp = − 4 μA/cm2; c An action potential induced by a stimulation and the afterpotential following the action potential
The Ih current is positive, then, the resting membrane potential prior to the Hopf bifurcation elevates with increasing gh. For instance, for Iapp = − 4 μA/cm2 (same as that used in the network), the resting membrane potential increases from − 73.16 mV to − 70.65 mV as gh increases from 0 (black) to 0.3 (blue) mS/cm2, as illustrated in Fig. 14b. The behavior before t = 50 ms of Fig. 14c is same as that of Fig. 14b, with y-axis shown in a large scale. Perturbed by a voltage pulse stimulation (V = 15 mV), an action potential is evoked for gh = 0 and 0.3 mS/cm2. The afterpotential for gh = 0.3 mS/cm2 is higher than that for 0 mS/cm2. For both gh = 0 and 0.3 mS/cm2, the afterpotential increases monotonically to recover to the resting membrane potential after a relatively long time (nearly 50 ms).
Response time curve to explain the conduction delay behavior
A current pulse with width of 2 ms and intensity ΔA of 8.5 μA/cm2 is applied at t = 0 ms, as shown in Fig. 15a1, resulting in the first action potentials for gh = 0 (black) and 0.3 (blue) mS/cm2. The level of the unrecovered afterpotential for gh = 0.3 mS/cm2 (blue) is higher than that for 0 mS/cm2 (black). Then, as the same pulse stimulus is applied to the unrecovered afterpotentials at a phase Δt (15 ms in Fig. 15a1) following the first pulse stimulation, the second action potentials are evoked. The application time/phase Δt corresponds to Ts, and ΔA corresponds to gc. The second action potential appears earlier for gh = 0.3 mS/cm2 (blue), compared with 0 mS/cm2 (black), showing that the response times of action potentials evoked at a same time are different for different gh values. The response time, which is defined as the interval between the beginning phase of the pulse stimulation and the peak of action potential evoked by the pulse, is 3.08 ms and 5.44 ms for gh = 0.3 and 0 mS/cm2, respectively.
The response time changes with respect to Δt, as illustrated in Fig. 15a2, black and blue for gh = 0 and 0.3 mS/cm2, respectively. The responses of the afterpotentials to a stimulation pulse with ΔA = 8.5 μA/cm2 (red) and ΔA = 9.5 μA/cm2 (blue) are illustrated in Fig. 15b1, b2, respectively, which can explain the conduction delay in three aspects:
For gh = 0 (black) and 0.3 (blue) mS/cm2, with increasing Δt, the response time decreases at first and then tends to be a fixed value, which remains unchanged as Δt increases further. The later the phase of current pulse stimulation applied to the afterpotential is, the faster the action potential is, resembling the changes of conduction time (delay) in the network with respect to Ts which determines the application phase of the coupling current pulse. Then, such a result can well explain that the conduction delay occurs for a shorter Ts than that of the faithful conduction behavior, and conduction time for the conduction delay behavior decreases with increasing Ts (Please refer to Fig. 4).
The response time for gh = 0.3 mS/cm2 (blue) is lower than that for 0 mS/cm2 (black), which means that the larger the gh is, the faster the response of an action potential is. With increasing gh, action potentials induced by the current pulse are more quickly, similar to the results induced by the increase of gh in the network. Such a result can well explain why the conduction time for gh = 0.3 mS/cm2 is shorter than that for 0 mS/cm2 (Please refer to Figs. 2, 3).
A smaller conduction time appears for a larger gc value (Please refer to Fig. 5) can be explained. The response time of an action potential for ΔA = 9.5 μA/cm2 (blue) becomes shorter, compared with ΔA = 8.5 μA/cm2 (red), as shown in Fig. 15b1. The response time is 3.43 ms for ΔA = 9.5 μA/cm2 and 5.44 ms for ΔA = 8.5 μA/cm2. The response time changes with respect to Δt, as shown in Fig. 15b2, red and blue for ΔA = 8.5 and ΔA = 9.5 μA/cm2, respectively. The response time for ΔA = 9.5 μA/cm2 is shorter, compared with ΔA = 8.5 μA/cm2, which means that the larger the current stimulation pulse strength (ΔA) is, the faster the action potential is. ΔA of the current stimulation pulse is associated with gc in the network to determine the coupling current strength.
Current threshold curve to explain the conduction failure behavior
A current pulse with application time Δt and stimulus intensity ΔA is applied to the afterpotential, as shown in Fig. 16a. The current pulse can evoke action potential or not, which can be used to explain the conduction delay/faithful conduction or conduction failure. For the afterpotential (gh = 0.3 mS/cm2), a suprathreshold current pulse (blue) can cause an action potential at the phase Δt, while a subthreshold current pulse (red) does not, as depicted in Fig. 16a. For Δt = 15 ms, pulse strength ΔA = 6 μA/cm2 (blue) is used as a suprathreshold stimulation, and ΔA = 4 μA/cm2 (red) as a subthreshold stimulation, as illustrated in Fig. 16a. The minimal ΔA to trigger an action potential at Δt is called the current threshold, denoted as AT(Δt). For different values of Δt, the values of the current threshold intensity AT(Δt) are different.
Fig. 16.
Current threshold curve to explain the conduction failure. a Responses of the afterpotentials to a current stimulus pulse with different intensities ΔA (blue and red curves for ΔA = 6 μA/cm2 and 4 μA/cm2, respectively). Pulse width 2 ms, application phase Δt = 15 ms, and gh = 0.3 mS/cm2; b Changes of the current intensity threshold AT with increasing ∆t for gh = 0 (black) and 0.3 (blue) mS/cm2; c The black and blue curves correspond to the border between conduction failure and conduction delay for gh = 0 in Fig. 8(a) and gh = 0.3 mS/cm2 in Fig. 8(b), respectively
The threshold AT(Δt) increases with decreasing Δt, for both gh = 0.3 (blue) and 0 (black) mS/cm2, as illustrated in Fig. 16b. The larger the gh value is, the smaller the AT(Δt) is. The current threshold curve decreases monotonically. The threshold curve can explain the conduction failure behavior in three aspects.
The increases of AT(Δt) with decreasing Δt for both gh = 0 and 0.3 mS/cm2, resembling changes of the border between the conduction failure and conduction delay behaviors in plane (Ts, gc), as shown in Fig. 16c, black and blue for gh = 0 and 0.3 mS/cm2, respectively. At a short Δt, AT(Δt) is very large. Therefore, the conduction failure mostly occurs at short Ts.
The current threshold AT(Δt) is relatively large in a wide range of Ts, showing that pulse stimulation strength ΔA larger than AT(Δt) can induce action potential while lower than AT(Δt) not, which can explain why the conduction failure occurs for smaller values of gc, compared with the conduction delay.
The threshold AT(Δt) becomes large with decreasing gh value, as illustrated in Fig. 16b, resembling that the border between the conduction delay and conduction failure (Fig. 8) becomes high as gh decreases, as depicted in Fig. 16c. The threshold AT(Δt) for gh = 0 mS/cm2 is higher, compared with 0.3 mS/cm2. Therefore, the failure rate for gh = 0 mS/cm2 is larger, compared with 0.3 mS/cm2. Then, the modulation effect that downregulation of gh enhances the conduction failure rate is explained.
Conclusion and discussion
Abnormal firings or spatiotemporal behaviors of the nervous system can lead to abnormal functions or brain diseases (Li et al. 2017; Yang et al. 2021; Song et al. 2022; Ma 2023; Wang et al. 2023). Conduction behavior of action potentials is one of the important spatiotemporal behaviors of the nervous system (Chen et al. 2019; Dai et al. 2022; Li et al. 2023). Different from the traditional view that action potentials faithfully conduct along axon to involve in normal function of the nervous system, non-faithful conduction behaviors such as the conduction failure or delay behaviors have been observed in the experiments associated with abnormal functions or diseases (Sun et al. 2012; Ballo et al. 2012; Zhang et al. 2017). For instance, conduction failure behavior is observed in C-fibers related to the pathological pain (Sun et al. 2012), and the conduction delay behavior is observed in the pyloric dilator axons to investigate the temporal fidelity of action potentials and the temporal encoding (Ballo et al. 2012). Furthermore, the influence of the Ih current in modulating conduction behaviors is investigated. Increase of the Ih current reduces conduction delay to improve the temporal encoding (Wang et al. 2016a; Zhang et al. 2017), and blockage of the Ih current significantly increases the conduction failure to ease pain (Wang et al. 2016a). In this article, the experimental observations of the conduction delay and failure behaviors are reproduced in a chain network model, and the dynamical mechanism for the modulations of the Ih current to the conduction delay and failure behaviors is presented with the nonlinear responses of afterpotentials in a neuronal model. The results exhibit significance in the following aspects:
Firstly, the conduction delay and failure behaviors closely matching the experimental results in 3 aspects are reproduced, and the direct relationships between conduction delay and failure behaviors are obtained. (1) Conduction failure and delay behaviors appear for action potentials with a high frequency, i.e., the period (Ts) of the stimulations to evoke action potentials is short. With increasing Ts, the conduction delay and failure reduce at first, and change to faithful conduction behavior at last. (2) Conduction failure and delay behaviors appear for the fiber with little conduction velocity, i.e., little coupling strength (gc), whereas faithful conduction behavior occurs for large gc or long Ts. With decreasing gc, the conduction delay and failure increase. Compared with the conduction delay behavior, the conduction failure behavior appears for smaller gc and shorter Ts, which presents direct relationships between the conduction delay and failure behaviors. In the previous experiments, no direct relationship between conduction delay and failure behaviors has been presented, since the two behaviors are studied independently in different types of fibers, as mentioned above (Sun et al. 2012; Ballo et al. 2012). In addition, for the conduction failure behavior, conduction time of action potentials increases before the action potential fails to conduct, as shown in Fig. 7c, which presents simulation results to the experimental observation (Zhu et al. 2009). (3) Upregulation of the Ih current reduces the conduction delay, and downregulation of the Ih current enhances the conduction failure rate. The enhancement of the failure rate with changing gh contains four cases of changes of spatiotemporal behaviors in wide ranges of the plane (Ts, gc). The result shows that upregulation of the Ih current can help to improve the temporal encoding and blockage of the Ih current can help to ease pain (less action potentials mean lower pain sensation).
Secondly, the response time and current threshold of an action potential evoked from the unrecovered afterpotentials by a current pulse stimulation are acquired in a neuron model, which can well explain the modulations of the coupling strength (gc) and frequency of action potentials (Ts) on the conduction delay and failure behaviors. Conduction delay and failure behaviors appear for action potentials with a high frequency for a short Ts, due to that the afterpotentials cannot recover to the resting potential within the short Ts. As a stimulation pulse with strength ΔA is introduced to the unrecovered afterpotentials at a phase Δt, an enhanced current threshold and a prolonged response time of an action potential appear, which present cause for the conduction failure and conduction delay, respectively. Δt and ΔA of the current pulse stimulation respectively correspond to Ts and gc. With increasing Δt, the current threshold and response time decrease, presenting the cause for the decrease of the conduction delay and failure with increasing Ts. The existence of a large threshold at a small Δt shows that stimulation larger than the threshold can evoke an action potential and lower than the threshold not, which can explain that the conduction failure occurs for a smaller gc, compared with the conduction delay. In the previous studies, conduction delay behavior is observed in experiments and reproduced in theoretical model (Ballo et al. 2012; Zhang et al. 2017). In this article, the conduction delay behavior is further explained with the response time of action potential. In the previous studies, the conduction failure behavior is reproduced in multiple ranges of short Ts values (Zhang and Gu 2019), however, such a discrete distribution of Ts has not been reported in the experimental studies. In this paper, the conduction failure behavior appears in a continuous range of Ts with short values, resembling the experimental observations (Zhu et al. 2009). The different Ts ranges for the conduction failure are due to that the current threshold curve exhibits damping oscillations with multiple cycles in Ref (Zhang and Gu 2019), whereas a continuous range of Ts in the present paper is due to that the current threshold curve decreases monotonically.
Finally, modulations of the Ih current on the conduction delay and failure behaviors are explained with the response time and current threshold. As the Ih current increases, Hopf bifurcation with respect to current advances, and the resting membrane potential before the Hopf bifurcation elevates. Then, the level of the afterpotentials elevates, resulting in shortened response time and reduced current threshold to cause an action potential. Then, the shortened response time can explain decrease of the conduction delay induced by upregulation of the Ih current observed in the experiments (Ballo et al. 2012; Zhang et al. 2017), which lays a theoretical foundation to improve temporal encoding by upregulation of the Ih current. If the Ih current decreases, the afterpotential becomes low, resulting in a prolonged response time and enhanced current threshold. The enhanced current threshold can explain the enhanced failure rate induced by downregulation of the Ih current observed in the experiments (Wang et al. 2016a; Byczkowicz et al. 2019), laying a theoretical foundation to ease pain through blockage of the Ih current. In previous studies, upregulation of K+ channel enhances the conduction failure (Zhang et al. 2020), showing another target to modulate the pain sensation. The results of this paper present a comprehensive viewpoint for the conduction delay and failure behaviors of action potentials with a high frequency and lay a theoretical foundation that the Ih current is an important target to modulate temporal encoding and pathological pain.
Although some progress has been made in the present paper, there are many questions related to the nonfaithful conduction behaviors or the modulation effects of the Ih current to be studied. Firstly, except for the temporal encoding and pathological pain (Wang et al. 2016a; Zhang et al. 2017), non-faithful conduction behaviors of action potentials such as conduction failure behavior are associated with a variety of pathological states or diseases such as epilepsy, Parkinson’s diseases, and multiple sclerosis (Waxman 2006; Lesnick et al. 2007; Wimmer et al. 2010). Then, the roles of the nonfaithful conduction in the brain diseases should be an important issue. Secondly, except for action potential frequency and ionic current (Cross and Robertson 2016; Feng et al. 2017; Cho et al. 2017), conduction behaviors of action potentials are related to many factors such as the fiber type or topology (Bucher and Goaillard 2011; Bukh et al. 2023), heterogeneity of the ion channel distribution (Freeman et al. 2016; Zang et al. 2022; Zang and Marder 2023), astrocyte (Lezmy et al. 2021), temperature (Song et al. 2019), and noise (Zhang et al. 2019b). The influence of other important factors on the nonfaithful conduction behaviors should be studied in future. For example, spontaneous firing induced by the uneven distribution of ion channel density along the axon is related to demyelinating diseases (Freeman et al. 2016), and slow inactivation of Na+ channel reduces firing frequency of axon to restore the conduction behavior (Zang et al. 2022). Finally, except for the conduction behaviors, the Ih current is also responsible for regulating the dynamics of the neuronal firing to achieve some important or specific functions (Zhang et al. 2019a; Guan et al. 2019, 2021; Takagi et al. 2020) such as sleep, learning, memory, sensation, and some brain diseases. Then, the various nonlinear dynamics related to the Ih current should be studied in future. More interestingly, in a recent study (Lezmy et al. 2021), the increase of Ca2+ concentration in astrocyte causes the release of ATP at first, and then induces the Ih current increased through a series of processes, resulting in the reduced conduction velocity of action potentials along axon with myelin, which shows that the modulation effect of the Ih current on the conduction behavior is in contrast to the results of the previous experiments (Ballo et al. 2012; Zhang et al. 2017) and the present paper. Then, the dynamics of the conduction behaviors of action potentials along axon with myelin, which is modulated by the Ih current and astrocyte, should be studied in future.
Acknowledgements
This work was sponsored by the National Natural Science Foundation of China (Grant Numbers: 12072236 and 12202147) and Nature Science Foundation of Henan province (Grant Number: 212300410196)
Data availability
The data used in this study are available upon request from the corresponding author.
Declarations
Conflict of interest
The authors indicate no conflicts of interest related to the content of this paper.
Footnotes
Publisher's Note
Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
References
- Arancibia-Cárcamo IL, Ford MC, Cossell L et al (2017) Node of Ranvier length as a potential regulator of myelinated axon conduction speed. Elife 6:e23329. 10.7554/eLife.23329 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ballo AW, Bucher D (2009) Complex intrinsic membrane properties and dopamine shape spiking activity in a motor axon. J Neurosci 29:5062–5074. 10.1523/JNEUROSCI.0716-09.2009 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ballo AW, Nadim F, Bucher D (2012) Dopamine modulation of Ih improves temporal fidelity of spike propagation in an unmyelinated axon. J Neurosci 32:5106–5119. 10.1523/JNEUROSCI.6320-11.2012 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bernard Healey SA, Scholtes I, Abrahams M et al (2021) Role of hyperpolarization-activated cyclic nucleotide-gated ion channels in neuropathic pain: a proof-of-concept study of ivabradine in patients with chronic peripheral neuropathic pain. Pain Reports 6:e967. 10.1097/PR9.0000000000000967 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bucher D, Goaillard JM (2011) Beyond faithful conduction: Short-term dynamics, neuromodulation, and long-term regulation of spike propagation in the axon. Prog Neurobiol 94:307–346. 10.1016/j.pneurobio.2011.06.001 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bukh AV, Shepelev IA, Elizarov EM et al (2023) Role of coupling delay in oscillatory activity in autonomous networks of excitable neurons with dissipation. Chaos 33:073114. 10.1063/5.0147883 [DOI] [PubMed] [Google Scholar]
- Byczkowicz N, Eshra A, Montanaro J et al (2019) HCN channel-mediated neuromodulation can control action potential velocity and fidelity in central axons. Elife 8:e42766. 10.7554/eLife.42766 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Cartwright R (2010) A guide to XPPAUT for researchers and students. Perspect Public Health 130:239–239. 10.1177/1757913910379198 [Google Scholar]
- Chaplan SR, Guo HQ, Lee DH et al (2003) Neuronal hyperpolarization-activated pacemaker channels drive neuropathic pain. J Neurosci 23:1169–1178. 10.1523/jneurosci.23-04-01169.2003 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chen H, Garcia-Gonzalez D, Jérusalem A (2019) Computational model of the mechanoelectrophysiological coupling in axons with application to neuromodulation. Phys Rev E 99:032406. 10.1103/PhysRevE.99.032406 [DOI] [PubMed] [Google Scholar]
- Cho IH, Panzera LC, Chin M, Hoppa MB (2017) Sodium channel β2 subunits prevent action potential propagation failures at axonal branch points. J Neurosci 37:9519–9533. 10.1523/JNEUROSCI.0891-17.2017 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Cross KP, Robertson RM (2016) Ionic mechanisms maintaining action potential conduction velocity at high firing frequencies in an unmyelinated axon. Physiol Rep 4:e12814. 10.14814/phy2.12814 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Crotty P, Levy WB (2007) Effects of Na+ channel inactivation kinetics on metabolic energy costs of action potentials. Neurocomputing 70:1652–1656. 10.1016/j.neucom.2006.10.063 [Google Scholar]
- Dai S, Lu L, Wei Z et al (2022) Influence of temperature and noise on the propagation of subthreshold signal in feedforward neural network. Chaos, Solitons Fractals 164:112762. 10.1016/j.chaos.2022.112762 [Google Scholar]
- Dilley A, Richards N, Pulman KG, Bove GM (2013) Disruption of fast axonal transport in the rat induces behavioral changes consistent with neuropathic pain. J Pain 14:1437–1449. 10.1016/j.jpain.2013.07.005 [DOI] [PubMed] [Google Scholar]
- Ding Q, Jia Y (2021) Effects of temperature and ion channel blocks on propagation of action potential in myelinated axons. Chaos 31:053102. 10.1063/5.0044874 [DOI] [PubMed] [Google Scholar]
- Duan L, Liang W, Ji W, Xi H (2020) Bifurcation patterns of bursting within pre-bötzinger complex and their control. Int J Bifurc Chaos 30:2050192. 10.1142/S0218127420501928 [Google Scholar]
- Dutta DJ, Woo DH, Lee PR et al (2018) Regulation of myelin structure and conduction velocity by perinodal astrocytes. Proc Natl Acad Sci U S A 115:11832–11837. 10.1073/pnas.1811013115 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Feng Z, Wang Z, Guo Z et al (2017) High frequency stimulation of afferent fibers generates asynchronous firing in the downstream neurons in hippocampus through partial block of axonal conduction. Brain Res 1661:67–78. 10.1016/j.brainres.2017.02.008 [DOI] [PubMed] [Google Scholar]
- Freeman WJ (2000) Characteristics of the synchronization of brain activity imposed by finite conduction velocities of axons. Int J Bifurc Chaos 10:2307–2322. 10.1142/S0218127400001547 [Google Scholar]
- Freeman SA, Desmazières A, Fricker D et al (2016) Mechanisms of sodium channel clustering and its influence on axonal impulse conduction. Cell Mol Life Sci 73:723–735. 10.1007/s00018-015-2081-1 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gillespie MJ, Stein RB (1983) The relationship between axon diameter, myelin thickness and conduction velocity during atrophy of mammalian peripheral nerves. Brain Res 259:41–56. 10.1016/0006-8993(83)91065-X [DOI] [PubMed] [Google Scholar]
- Guan L, Jia B, Gu H (2019) A novel threshold across which the negative stimulation evokes action potential near a saddle-node bifurcation in a neuronal model with Ih current. Int J Bifurc Chaos 29:1950198. 10.1142/S0218127419501980 [Google Scholar]
- Guan L, Gu H, Zhao Z (2021) Dynamics of subthreshold and suprathreshold resonance modulated by hyperpolarization-activated cation current in a bursting neuron. Nonlinear Dyn 104:577–601. 10.1007/s11071-021-06230-8 [Google Scholar]
- Guo Y, Zhou P, Yao Z, Ma J (2021) Biophysical mechanism of signal encoding in an auditory neuron. Nonlinear Dyn 105:3603–3614. 10.1007/s11071-021-06770-z [Google Scholar]
- Hamada MS, Popovic MA, Kole MHP (2017) Loss of saltation and presynaptic action potential failure in demyelinated axons. Front Cell Neurosci 11:45. 10.3389/fncel.2017.00045 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hodgkin AL (1954) A note on conduction velocity. J Physiol 125:221–224. 10.1113/jphysiol.1954.sp005152 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hodgkin AL, Huxley AF (1952) A quantitative description of membrane current and its application to conduction and excitation in nerve. J Physiol 117:500–544. 10.1113/jphysiol.1952.sp004764 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Jansen LAR, Forster LA, Smith XL et al (2021) Changes in peripheral HCN2 channels during persistent inflammation. Channels 15:165–179. 10.1080/19336950.2020.1870086 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Khurana S, Remme MWH, Rinzel J, Golding NL (2011) Dynamic interaction of Ih and IK-LVA during trains of synaptic potentials in principal neurons of the medial superior olive. J Neurosci 31:8936–8947. 10.1523/JNEUROSCI.1079-11.2011 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kiziltan E, Dalkilic N, Guney FB, Pehlivan F (2007) Conduction velocity distribution: Early diagnostic tool for peripheral neuropathies. Int J Neurosci 117:203–213. 10.1080/00207450600582496 [DOI] [PubMed] [Google Scholar]
- Kueh D, Barnett WH, Cymbalyuk GS, Calabrese RL (2016) Na+/K+ pump interacts with the Ih-current to control bursting activity in central pattern generator neurons of leeches. Elife 5:e19322. 10.7554/eLife.19322 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lesnick TG, Papapetropoulos S, Mash DC et al (2007) A genomic pathway approach to a complex disease: axon guidance and parkinson disease. PLoS Genet 3:0984–0995. 10.1371/journal.pgen.0030098 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lezmy J, Arancibia-Cárcamo IL, Quintela-López T et al (2021) Astrocyte Ca2+-evoked ATP release regulates myelinated axon excitability and conduction speed. Science 374:eabh2858. 10.1126/science.abh2858 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Li JJ, Xie Y, Yu YG et al (2017) A neglected GABAergic astrocyte: Calcium dynamics and involvement in seizure activity. Sci China Technol Sci 60:1003–1010. 10.1007/s11431-016-9056-2 [Google Scholar]
- Li T, Wu Y, Yang L et al (2023) Neuronal morphology and network properties modulate signal propagation in multi-layer feedforward network. Chaos, Solitons Fractals 172:113554. 10.1016/j.chaos.2023.113554 [Google Scholar]
- Lyu W, Li S, Chen Z, Bi Q (2023) Bursting dynamics in a singular vector field with codimension three triple zero bifurcation. Mathematics 11:2486. 10.3390/math11112486 [Google Scholar]
- Ma J (2023) Biophysical neurons, energy, and synapse controllability: a review. J Zhejiang Univ Sci A 24:109–129. 10.1631/jzus.A2200469 [Google Scholar]
- Mao H, Wang X, Chen W et al (2019) Use of in vivo single-fiber recording and intact dorsal root ganglion with attached sciatic nerve to examine the mechanism of conduction failure. J vis Exp 150:e59234. 10.3791/59234 [DOI] [PubMed] [Google Scholar]
- Panzeri S, Brunel N, Logothetis NK, Kayser C (2010) Sensory neural codes using multiplexed temporal scales. Trends Neurosci 33:111–120. 10.1016/j.tins.2009.12.001 [DOI] [PubMed] [Google Scholar]
- Ratas I, Pyragas K (2012) Effect of high-frequency stimulation on nerve pulse propagation in the FitzHugh-Nagumo model. Nonlinear Dyn 67:2899–2908. 10.1007/s11071-011-0197-x [DOI] [PubMed] [Google Scholar]
- Roth FC, Hu H (2020) An axon-specific expression of HCN channels catalyzes fast action potential signaling in GABAergic interneurons. Nat Commun 11:2248. 10.1038/s41467-020-15791-y [DOI] [PMC free article] [PubMed] [Google Scholar]
- Shimba K, Kotani K, Jimbo Y (2021) Microfabricated device to record axonal conduction under pharmacological treatment for functional evaluation of axon ion channel. IEEE Trans Biomed Eng 68:3574–3581. 10.1109/TBME.2021.3078473 [DOI] [PubMed] [Google Scholar]
- Song X, Wang H, Chen Y, Lai Y (2019) Emergence of an optimal temperature in action-potential propagation through myelinated axons. Phys Rev E 100:032416. 10.1103/PhysRevE.100.032416 [DOI] [PubMed] [Google Scholar]
- Song J, Liu S, Lin H (2022) Model-based quantitative optimization of deep brain stimulation and prediction of Parkinson’s states. Neuroscience 498:105–124. 10.1016/j.neuroscience.2022.05.019 [DOI] [PubMed] [Google Scholar]
- Sun W, Miao B, Wang XC et al (2012) Reduced conduction failure of the main axon of polymodal nociceptive C-fibres contributes to painful diabetic neuropathy in rats. Brain 135:359–375. 10.1093/brain/awr345 [DOI] [PubMed] [Google Scholar]
- Takagi D, Okamoto Y, Ohba T et al (2020) Comparative study of hyperpolarization-activated currents in pulmonary vein cardiomyocytes isolated from rat, guinea pig, and rabbit. J Physiol Sci 70:6. 10.1186/s12576-020-00736-3 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wang X, Wang S, Wang W et al (2016a) A novel intrinsic analgesic mechanism: the enhancement of the conduction failure along polymodal nociceptive C-fibers. Pain 157:2235–2247. 10.1097/j.pain.0000000000000632 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wang X, Wang S, Zhang M et al (2016b) Α-dendrotoxin-sensitive Kv1 channels contribute to conduction failure of polymodal nociceptive C-fibers from rat coccygeal nerve. J Neurophysiol 115:947–957. 10.1152/jn.00786.2014 [DOI] [PubMed] [Google Scholar]
- Wang X, Yu Y, Han F, Wang Q (2023) Dynamical mechanism of parkinsonian beta oscillation in a heterogenous subthalamopallidal network. Nonlinear Dyn 111:10505–10527. 10.1007/s11071-023-08381-2 [Google Scholar]
- Waxman SG (1977) Conduction in myelinated, unmyelinated, and demyelinated fibers. Arch Neurol 34:585–589. 10.1001/archneur.1977.00500220019003 [DOI] [PubMed] [Google Scholar]
- Waxman SG (2006) Axonal conduction and injury in multiple sclerosis: The role of sodium channels. Nat Rev Neurosci 7:932–941. 10.1038/nrn2023 [DOI] [PubMed] [Google Scholar]
- Weng X, Smith T, Sathish J, Djouhri L (2012) Chronic inflammatory pain is associated with increased excitability and hyperpolarization-activated current (Ih) in C- but not Aδ-nociceptors. Pain 153:900–914. 10.1016/j.pain.2012.01.019 [DOI] [PubMed] [Google Scholar]
- Wimmer VC, Reid CA, So EYW et al (2010) Axon initial segment dysfunction in epilepsy. J Physiol 588:1829–1840. 10.1113/jphysiol.2010.188417 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Xing M, Yang Z, Chen Y (2023) Bursting types and bifurcation analysis of the temperature-sensitive Purkinje neuron. Nonlinear Dyn 111:1819–1834. 10.1007/s11071-022-07917-2 [Google Scholar]
- Yang C, Liu Z, Wang Q et al (2021) Epileptic seizures in a heterogeneous excitatory network with short-term plasticity. Cogn Neurodyn 15:43–51. 10.1007/s11571-020-09582-w [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yao C, Yao Y, Qian Y, Xu X (2022) Temperature-controlled propagation of spikes in neuronal networks. Chaos, Solitons Fractals 164:112667. 10.1016/j.chaos.2022.112667 [Google Scholar]
- Zang Y, Marder E (2021) Interactions among diameter, myelination, and the Na/K pump affect axonal resilience to high-frequency spiking. Proc Natl Acad Sci U S A 118:e2105795118. 10.1073/pnas.2105795118 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zang Y, Marder E (2023) Neuronal morphology enhances robustness to perturbations of channel densities. Proc Natl Acad Sci 120:e2219049120. 10.1073/pnas.2219049120 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zang Y, Marder E, Marom S (2022) Sodium channel slow inactivation normalizes firing in axons with uneven conductance distributions. BioRxiv 33:1818–1824. 10.1016/j.cub.2023.03.043 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhang X, Gu H (2019) Influence of hopf bifurcation dynamics on conduction failure of action potentials along nerve C-fiber. Int J Bifurc Chaos 29:1950093. 10.1142/S0218127419500937 [Google Scholar]
- Zhang Y, Bucher D, Nadim F (2017) Ionic mechanisms underlying history-dependence of conduction delay in an unmyelinated axon. Elife 6:e25382. 10.7554/eLife.25382 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhang M, Guo H, Ma Y et al (2019a) Acupoint sensitization is associated with increased excitability and hyperpolarization-activated current (Ih) in C- but not Aδ-type neurons. Neuroscience 404:499–509. 10.1016/j.neuroscience.2019.02.028 [DOI] [PubMed] [Google Scholar]
- Zhang X, Gu H, Guan L (2019b) Stochastic dynamics of conduction failure of action potential along nerve fiber with Hopf bifurcation. Sci China Technol Sci 62:1502–1511. 10.1007/s11431-018-9515-4 [Google Scholar]
- Zhang X, Gu H, Wu F (2019c) Memristor reduces conduction failure of action potentials along axon with Hopf bifurcation. Eur Phys J Spec Top 228:2053–2063. 10.1140/epjst/e2019-900004-2 [Google Scholar]
- Zhang X, Gu H, Ma K (2020) Dynamical mechanism for conduction failure behavior of action potentials related to pain information transmission. Neurocomputing 387:293–308. 10.1016/j.neucom.2019.12.114 [Google Scholar]
- Zhu ZR, Tang XW, Wang WT et al (2009) Conduction failures in rabbit saphenous nerve unmyelinated fibers. Neurosignals 17:181–195. 10.1159/000209279 [DOI] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
Data Availability Statement
The data used in this study are available upon request from the corresponding author.














