Spike solutions in Gierer-Meinhardt model with a time dependent anomaly exponent 1 2 Yana Nec Thompson Rivers University, 805 TRU Way, Kamloops, British Columbia, Canada, 3 4 email: cranberryana@gmail.com 5 Abstract 6 Experimental evidence of complex dispersion regimes in natural systems, where the growth of the mean square displacement in time cannot be characterised by a single power, has been accruing for the past two decades. In such processes the exponent γ(t) in hr2 i ∼ tγ(t) at times might be approximated by a piecewise constant function, or it can be a continuous function. Variable order differential equations are an emerging mathematical tool with a strong potential to model these systems. However, variable order differential equations are not tractable by the classic differential equations theory. This contribution illustrates how a classic method can be adapted to gain insight into a system of this type. Herein a variable order Gierer-Meinhardt model is posed, a generic reaction– diffusion system of a chemical origin. With a fixed order this system possesses a solution in the form of a constellation of arbitrarily situated localised pulses, when the components’ diffusivity ratio is asymptotically small. The pattern was shown to exist subject to multiple step-like transitions between normal diffusion and sub-diffusion, as well as between distinct sub-diffusive regimes. The analytical approximation obtained permits qualitative analysis of the impact thereof. Numerical solution for typical cross-over scenarios revealed such features as earlier equilibration and non-monotonic excursions before attainment of equilibrium. The method is general and allows for an approximate numerical solution with any reasonably behaved γ(t). Keywords: fractional differential equations, matched asymptotic expansions, variable order differential equations, numerical estimates of memory integrals 1. Background 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 Beginning circa 1980 measurements of dispersion in various natural systems documented behaviour that could not be captured faithfully by a linear growth of the mean square displacement hr2 i ∼ t. In time such systems were termed anomalous. Spanning numerous scientific fields, the underpinning diffusive processes can be catalogued by basic transport properties, often including the Fickian diffusion as a special limit (Codling et al., 2008; Eliazar and Klafter, 2011). The sui generis nature of each anomalous process notwithstanding, wide sub-classes have been identified (Klages et al., 2008). The present contribution addresses a type of anomaly named sub-diffusion due to the sub-linear dispersion hr2 i ∼ tγ with 0 < γ < 1. One way to obtain this kind of transport is to generalise the integer derivative in the diffusion equation to one of a fractional order (Metzler and Klafter, 2000). Mathematically an integer Preprint submitted to Elsevier 7 8 June 12, 2017 24 25 26 27 28 29 30 31 32 33 2 34 35 order partial differential equation for the concentration of components u with a constant diffusion coefficient D and source f (u) ut = D∆u + f (u), 36 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 61 62 63 64 65 66 67 68 69 70 t > 0, (1a) t > 0. (1b) will become a fractional partial differential equation uγt = D∆u + f (u), 37 r ∈ Ω, r ∈ Ω, Over a certain range of time the dispersion r = |r| in the domain Ω will exhibit sub-diffusion with a mean square displacement that grows in time with the exponent γ. Within this framework the index γ need not be constant, albeit traditionally it has been fixed. Whilst the integer derivative operator is indisputably unique, multifarious fractional derivatives exist. For a constant γ the most classic definition appears in Oldham and Spanier (1974), whereas more specialised operators can be found for instance in Elliot (1993) and Chen et al. (2010), as well as in studies of specific systems quoted below. Variable order concominants have drawn some attention as well, cf. Naber (2004) and Ramirez and Coimbra (2009). Recent studies bespeak the potential purport of the latter operators as a modelling tool for systems evincing a time dependent dispersion power hr2 i ∼ tγ(t) . One group of applications manifests distinct scales γ at well defined periods of time, i.e. a piecewise constant approximation of γ(t) is a practical approach. An example thereof is a time limited process in a finite domain, as undergone by molecules in biological media (Reynolds, 2005; Bakalis et al., 2015), often with an auxiliary signalling reaction triggering an essential change in the spatial structure of the diffusing molecules or local medium (Cabal et al., 2006; Torreno-Pina et al., 2016). McKinley et al. (2009) construct a flexible model in the field of rheology, and Meerschaert et al. (2013) present a more formal framework. Another group of applications employs a continuous exponent γ(t). Sun et al. (2009) give examples of numerical solutions with a linear function in t (as well as exponents depending on variables other than time). Saxton (1997) examines the distribution of diffusion coefficients in single particle tracking and demonstrates how statistically the apparent mean square displacement hr2 i can be highly non-linear. The same author discusses a smooth cross-over between two regimes with constant γ and compares to non-transient diffusion, connecting the results to experimental measurement techniques (Saxton, 2007). At times γ and the conforming order of the diffusion equation are noted to depend on system variables such as concentration (Chen et al., 2013) or temperature (Morgan and Spera, 2001), however ultimately when the mean square displacement is measured, the exponent γ will be timedependent. In light of the above, variable order fractional equations promise a significant advance in questions of correct interpretation of experimental measurements and juxtaposition of various particle tracking techniques. To date studies presented numerical solutions that could be compared to classic cases or fitted to experimental results, but offered little analytical insight. An analytical solution to a paradigm system, explicating the temporal evolution subject to transitions between disparate diffusive regimes and possible new behaviour thereupon, will 3 be a valuable step. This study focusses on a phenomenological reaction – diffusion system with an aim to relax the restriction of a constant order temporal derivative (mean square displacement growth power), thereby permitting the description of cross-over from regular diffusion to sub-diffusion and between sub-diffusive regimes. 2. Gierer-Meinhardt model ∂tγ a = ǫ2γ axx − a + ap hq − 1 < x < 1, am τo ∂tγ h = hxx − h + ǫ−γ s h ax (±1, t) = hx (±1, t) = 0 , t > 0, 73 74 76 77 78 79 80 81 82 83 84 85 86 (2a) 87 − 1 < x < 1, t > 0, (2b) h(x, 0) = h0 (x) , (2c) 88 a(x, 0) = a0 (x) , where a(x, t) and h(x, t) are the activator and inhibitor concentrations, respectively. Here 0 < ǫ ≪ 1, τo > 0, the quadruple of reaction exponents (p, q, m, s) satisfies q > 0, 72 75 The reaction – diffusion system in question was first proposed by Gierer and Meinhardt (1972) and describes the interaction of two chemical species: an activator component that encourages the reaction process and an inhibitor, suppressing the rate thereof. When the diffusivity ratio between the components is asymptotically small with the reaction terms appropriately scaled, the system possesses a solution in the form of a set of spikes, i.e. localised spots of high concentration. With a fixed index γ these solutions were obtained by Iron et al. (2001) for regular diffusion, γ = 1, and by Nec and Ward (2012) for sub-diffusion, 0 < γ < 1. Here it is proposed to endow the problem with time variability of the anomaly index γ that is also the order of the partial differential equations, but in a way that still admits of a solution. On a finite, one-dimensional domain with no flux boundaries the unified model reads p > 1, 71 m > 0, s ⩾ 0, p−1 m < , q s+1 89 90 (2d) and the index γ ranges 0 < γ ⩽ 1. When γ = 1, system (2) comprises two conventional partial differential equations. When 0 < γ < 1, the derivative of order γ is defined (for a univariate function) as Z t 1 f (t) − f (t − ζ) dγ f (t) = − dζ, 0 < γ < 1, (2e) γ dt Γ(−γ) 0 ζ γ+1 wherein Γ denotes the Gamma function. Further detail on this operator can be found in Elliot (1993) and Nec and Ward (2012). Since the integration is with respect to ζ, it is forthwith admissible to extend the order to be a time dependent function γ(t). Some forms of memory operators include differentiation with respect to t or imbed the delay directly into the order through the dependence γ(t − ζ), therefore requiring a more careful generalisation, 91 92 93 94 95 96 97 98 4 108 cf. Ramirez and Coimbra (2009). Henceforth the anomaly index is deemed to be a function of time, designated thus explicitly γ(t) or merely γ for simplicity. For a constant γ the spike solution is constructed by the method of matched asymptotic expansions (Iron et al., 2001; Nec and Ward, 2012). On an infinite domain a similar construction is possible (Nec, 2016b). One of the tacit assumptions of this classic method is that the partial differential equations are of a fixed order, as are the concomitant asymptotic scales, together admitting a reduction to a system of non-linear algebraic equations for the spikes’ heights and ordinary differential equations for the drift of loci. As shown hereinafter however, this reduction can be extended to the case of multiple transitions between a sequence of arbitrary orders within the range 0 < γ ⩽ 1 at equally arbitrary cross-over moments. 109 3. A pattern of n spikes 99 100 101 102 103 104 105 106 107 110 111 112 Consider an n-tuple of spikes centred at a set of arbitrary loci xi , {i = 0, . . . , n − 1}. Classically the outer solution for the activator away from the loci xi is the trivial quiescent state a(x, t) ≡ 0, whereas for the inhibitor h(x, t) expand h ∼ h(0) (x, t) + o(1). (3) . Within a narrow region about the locus the inner spatial variable yi = (x − xi ) ǫγ(t) is required to make the differential term on the right hand side in (2a) of order O(1). The corresponding slow time variable τ = ǫα(t) t. The temporal scale α is determined below. The inner asymptotic solutions are set as   (0) (1) γ(t) −α(t) Ai (yi , τ ) = a xi + ǫ yi , ǫ τ ∼ Ai (yi , τ ) + ǫγ(τ ) Ai (yi , τ ) + · · · (4a) def 113 114 115 116 117 118 119 120 121   (1) (0) Hi (yi , τ ) = h xi + ǫγ(t) yi , ǫ−α(t) τ ∼ Hi (yi , τ ) + ǫγ(τ ) Hi (yi , τ ) + · · · . (4b) The unusual feature about (4) is the time dependence of the scales γ and α, as conventionally these are constant. It is imperative to substantiate when an expansion of this ilk is valid. Obviously, it is valid when γ and α are constant. Observe that it is also valid for any step-like function: X n  o γ(t) = γj θ t − t× j−1 − θ t − t× j , j = 1, 2, . . . , (5) j 122 123 124 125 126 127 128 129 wherein 0 < γj ⩽ 1 are constant anomaly indices, θ(t) denotes the Heaviside step function, t× j are the corresponding cross-over times satisfying t× j > t× j−1 and t× 0 = 0 without loss of generality. With γ(t) ≡ 1 the diffusion mechanism is the regular Fickian diffusion entailing a linear growth of the mean square displacement in time hr2 i ∼ t, recovering the classic model (Gierer and Meinhardt, 1972). With 0 < γ(t) < 1 the dispersion is sub-diffusive, i.e. hr2 i ∼ tγ(t) . A plain, fixed order process is obtained with all γj equal. A regular to anomalous transition is given by γj = 1, 0 < γj+1 < 1 and vice versa, whilst 0 < γj < 1, γj 6= γj+1 yield transitions between distinct sub-diffusive regimes. At any cross-over moment 5 t× j the scales of the entire asymptotic series (4) will be instantaneously adjusted. Albeit the non-linearity introduced into series (4) through the dependence ǫγ(t) forthwith precludes an extension to functions of a more generic nature, it is hereby conjectured that the spike solution will persist (albeit not be described by the series above) for smooth functions such that the changes in γ(t) occur on a time scale much faster than τ . For instance for an inner layer transition γ(t) = t − t× j γj+1 + γj γj+1 − γj tanh + 2 ε 2 n−1 X (0) βm−s Hj 131 132 133 134 135 (6) this requirement is mathematically expressed as ε ≪ ǫ−α , posing virtually no restriction on the inner layer width ε since ǫ ≪ 1 and α > 0, whilst an inner layer exists as such for ε ∼ O(1) or ε ∼ o(1). Thence illatively follows that the functional form of γ(t) is encumbered but little, in particular any temporal variation on O(1) scale will not interfere with the formation of the spike and its slow drift. For the solution construction to hold formally it must be assumed that γ(t) is a piecewise constant function as in (5). However, any reasonably behaved function γ(t) might be discretised into a set of step-like segments. The combination of solutions will remain valid as long as t ∼ O(1) or higher, i.e. no attempt is made to seek a truly smooth evolution, or to express that more technically, the orders of magnitude are preserved. Small steps in γ are permissible as long as the restriction on t is observed. Figure 1 shows a possible discretisation to approximate the smooth function (6). For any given segment the classic reduction yields the following algebraic-differential (0) system (consult Nec and Ward (2012) for technical detail). The spikes’ heights Hi (leading order inner solution for the inhibitor) are governed by a set of non-linear equations (0) Hi (τ ) = −bm 130 G(xi ; xj ) , 136 137 138 139 140 141 142 143 144 145 146 147 148 149 150 (7a) j=0 where G(x, xi ) is the Green’s function   − csch 2 cosh(xi − 1) cosh(x + 1) G(x; xi ) =  − csch 2 cosh(xi + 1) cosh(x − 1) and bm = Z 151 x < xi (7b) x > xi 152 um dy , (7c) R is a factor computed from the homoclinic u(y) u(y) = p+1 2 !1/(p−1) sech 153 2/(p−1) p−1 y 2 ! (7d) 6 γ(t) γj γj+1 t× j t× j+1 t Figure 1: A valid discretisation example of a smooth cross-over. 154 155 156 157 that controls the spike’s shape. In particular the leading order inner solution for the activator is given by q (0) β (0) β= , (7e) Ai = Hi u(yi ), p−1 whence Ai (yi , τ ) decays exponentially as |yi | −→ ∞ and is localised in the vicinity of the spike locus xi . The leading order outer solution for the inhibitor reads (0) h (x, t) = −bm n−1 X (0) βm−s Hi G(x; xi ) . (7f) i=0 158 159 The spikes’ loci xi (τ ) drift on the slow time scale τ = ǫα t, α = γ + 1, according to the differential system       n−1 γ−1   X βm−s qbm f (p; γ) 1 (0) βm−s  dxi dxi (0) − + G (x ; x ) , = G (x ; x ) + G (x ; x ) + H H x i j x i i x i i j (0)  dτ dτ 2 i (p + 1)Hi    j=0   j6=i (7g) 7 160   f p; γ(t) ≡ Z 1 dxi Dγy u(y) ≡ sgn Γ(−γ) dτ up+1 dy R Z ∞( 0 ,Z u′ (y)Dγy u dy , (7h) R dxi u(y) − u y + ξ sgn dτ 161 !) dξ . ξ γ+1 (7i) With regular diffusion f (p; 1) = 2(p + 1)/(p − 1), equation (7h) being a fractional generalisation thereof. Operator (7i) ostensibly distinguishes between leftward and rightward drift, thus breaking the symmetry of motion on either side of an equilibrium point. This is an explicit aftermath of the memory inherent to all temporal fractional derivatives. Nonetheless no actual asymmetry ensues, since the operator’s only appearance is in the second integral in (7h), acting on the with the odd function u′ (y), so that  even function u(y), in conjunction the factor f p; γ(t) does not depend on sgn x′i (τ ). The function Dγy u is neither even nor odd for 0 < γ < 1 (consult figure 2). The functional dependence of f (p; γ) is given in figure 3. At a cross-over time t× j the value will be instantaneously adjusted by a jump between respective values γj and γj+1 on the curve corresponding to the same value of the kinetic exponent p. Bar the small interval of non-monotonicity for p = 2 (uppermost curve in figure 3), diminution in γ will always entail an increase in f (p; γ). Albeit classically the foregoing asymptotic reduction (7) was devised for a constant γ, a painstaking checking verifies its validity for a piecewise constant function γ(t). This result is not altogether trivial, because whilst the inner spatial scale involves the compound ǫγ , the slow time scale involves ǫα with α = γ + 1, whence a jump in γ cannot be forthwith absorbed in ǫ. By and large system (7) permits a qualitative understanding of a variable order dynamical system of a fractional order γ(t): within the limitations of the asymptotic theory unifying an arbitrary sequence of piecewise constant segments of γ(t), it shows analytically that the variable order will not have an overwhelming impact on the nature of the solutions. Given that from a mathematical point of vantage relaxing the constraint of a fixed order is devastating to the application of traditional solution and analysis tools, and at the outset might lead to unpredictable results, this insight supports the physical soundness of variable order models. Past computational works showed examples of this inference, e.g. Sun et al. (2009); Chen et al. (2013), however with the Gierer-Meinhardt model a more formal substantiation is possible. Moreover, owing to the generic nature of the piecewise constant discretisation, it is safe to surmise that for most systems possessing asymptotic solutions of any kind the same procedure will conform to a valid description for an arbitrary piecewise constant sequence of orders. In addition it is illustrated hereinafter for specific spike constellations that a discretisation within the delineated limits of the presented theory yields a phase plane trajectory remarkably smooth visually, i.e. a piecewise constant sequence provides a good approximation without exacting the toll of a full numerical simulation. This last feature pertains solely to the Gierer-Meinhardt model, of course, and cannot be extended illatively to other dynamical systems. System (7) is expected to be accurate by comparison to the full system (2) up to the 162 163 164 165 166 167 168 169 170 171 172 173 174 175 176 177 178 179 180 181 182 183 184 185 186 187 188 189 190 191 192 193 194 195 196 8 1.5 Dγy u x′i <0 0 −1 −y∞ y∞ 0 y Figure 2: Function Dγy u with the fractional operator Dγy defined in (7i) for equally spaced 0.1 ⩽ γ ⩽ 0.9 (solid curves, neither dxi even nor odd functions) generalising u′ (dashed curve, odd function). The computation was performed for < 0 and dτ p = 2 upon regularisation of (2e) via integration by parts twice. The infinite bound of the integral was truncated at 3y∞ . Qualitatively similar results were obtained for 2 ⩽ p ⩽ 5. Cf. figure 1 of Nec and Ward (2012), computed to higher accuracy here for expositional purposes. 197 198 199 200 201 202 203 204 205 206 207 208 209 210   order of the correction term in series (4), i.e. O ǫγ(τ ) . The concomitant error with regular diffusion is O(ǫ). Similar reduced reaction– diffusion systems with regular diffusion have been shown to manifest excellent agreement with the full partial differential system, insofar as to predict correctly the onset of instability, cf. Sun et al. (2005) and Tzou et al. (2011). In stark contrast to integer order equations, partial differential equations of a fractional order, even without singular perturbations, require custom devised schemes, since conventional schemes have proved to baffle the regular notions as regards consistence, stability and accuracy (Meerschaert and Tadjeran, 2004). These difficulties are caused by the presence of memory, having at times unpredictable interference with other terms in the equations to be solved. In addition, time fractional operators contain integrals with singular kernels. Without exception successfully implemented schemes circumvented this fundamental quandary by regularising the operator. The modification involved either a change in the kernel, such as Caputo’s derivative, several examples of which appear in Ramirez and Coimbra (2009), or addition of an initial term (Henry and Wearne, 2000) or both (Sun et al., 2009). Defi- 9 f (p; γ) 9 6 4 3 0 1 γ 9 1 ⩽ γ ⩽ 10 . Cross marks show the normal values Figure 3: Factor f (p; γ) for equally spaced 2 ⩽ p ⩽ 5 computed for 10 + 2(p + 1)/(p − 1). f (p; γ) −→ ∞ at the stagnation limit γ −→ 0 . Cf. figure 3 of Nec and Ward (2012), computed to higher accuracy here for expositional purposes. nition (2e) is not regularised, its particular form essential to preserve basic equilibrium and spectrum related properties of (2) as a dynamical system (more detail can be found in Nec (2016b)). Thus a numerical solution of (2) is impossible with standard computational means. Nonetheless, the error bound above can be used to calculate the maximal value of ǫ that would ensure the solution given by (7) not exceed the corresponding error in the normal system by taking min γ(τ ). 4. Quasi-equilibrium solutions 212 213 214 215 216 217 System (7) is the reduction of (2) at the limit ǫ −→ 0. Different constellations of spikes were investigated to elucidate the effect of anomaly index transition on the drift toward equilibrium. Equation (7g) is a somewhat unconventionally written ordinary differential equation that can be cast into a better amenable to numerical integration form by x′i (τ ) x′i (τ ) 211 γ−1 = 218 219 220 221 γ sgn x′i (τ ) x′i (τ ) , and thence x′i (τ ) can be computed by evaluating the right hand side of (7g), whose sign determines sgn x′i (τ ). Generally at every integration step the non-linear system (7a) must be solved, albeit some simple patterns permit an explicit solution. In 222 223 224 10 225 226 227 228 229 230 231 232 233 234 235 236 237 238 239 240 241 242 243 244 245 246 247 the classic Runge-Kutta procedure employed here or other predictor-corrector methods (7a) must be solved for all intermediate predictor steps as well. A more fundamental peculiarity of (7g) is that the integration variable τ relates to the physical time t through τ = tǫγ(t)+1 , therefore changing scale at every cross-over point t× j , prescribed by equation (5). In praxis without lost of generality every  characterised  interval (0) – to the values by a fixed value γj was integrated with τ initialised to zero and xi , Hi reached at the end of the preceding interval. The full variable τ was made monotonic upon completion of integration by addition of τ× j = t× j ǫγj +1 . Infra solutions are presented for two groups of transition scenarios. The first category examines a sequence of significant jumps in γ with the purpose to exemplify the extent of abruptness in the pattern’s response with respect to relative velocity of the spikes and their heights; demonstrate the variability of possible response behaviours; present non-intuitive or non-monotonic response to certain parameters; identify typical mechanisms responsible for pattern evolution that might be considered advantageous for a natural system, such as a capability of fast equilibration or quick non-linear excursions before attainment of equilibrium. The second category pertains to the smooth cross-over. Function (6) was used with the corresponding piecewise constant sequence in figure 1. Despite the coarseness of this discretisation all evolution trajectories appeared absolutely smooth. Thus only a concise sample is shown depicting this observation. 4.1. Single spike With n = 1 equation (7a) is forthwith soluble for the spike height as a function of the locus location x0  −1/(βm−s−1) (0) H0 = − bm G(x0 , x0 ) , (8a) whereby the differential equation (7g) can be integrated in closed form for γ = 1 to yield    cosh(2x ) − 1 cosh(2x (0)) + 1 0 0 −4qτ sinh(2x0 ) 1   = ln + cosh(2) ln  (8b) p−1 sinh(2x0 (0)) 2 cosh(2x ) + 1 cosh(2x (0)) − 1 0 248 249 250 251 252 253 0 or numerically for 0 < γ < 1. Moreover, for any γ the sole fixed point x0 = 0 is stable, allowing to replace sgn x′i (τ ) = − sgn xi . From (7g) it is readily inferred that the only effect of γ is the speed wherewith the spike approaches equilibrium. What is less immediate is whether a change in γ incurs magnification or diminution of said speed. The magnitude of the right hand side of (7g) is below unity for any practical values of parameters, exclude for instance extremely small γ since lim+ f (p; γ) = ∞. Thereby x′i (τ ) will involve this γ−→0 254 255 256 257 258 quantity’s 1/γ > 1 power for any 0 < γ < 1. Bar a narrow region just below γ = 1 for p = 2, f (p; γ) is a descending function of γ (figure 3). Therefore the 1/γ > 1 power and growth of f (p; γ) are opposing influences, begetting the following behaviour: for a small increase in f , i.e. γ slightly below unity, the spike will slow down somewhat, for more significant growth of f it will speed up, whilst for small γ its progress will again be delayed at the cross-over point 11 t× , this latter evolution occurring over a very narrow interval of γ. This is best exemplified when f depends monotonically on γ, i.e. p > 2, although it is always correct. Figure 4 shows a typical example of this phenomenon. 259 260 x′0 (τ ) max x′0 0 0 τ× τ Figure 4: Anomaly dependent acceleration and delay phenomenon during the drift of a single spike: velocity x′0 (τ ) for a transition scenario γ(t < t× ) = γ1 , γ(t > t× ) = γ2 with γ1 = 1 and γ2 = 1 (dotted), γ2 = 12 (dash-dotted), γ2 = 0.0995 (dashed) and γ2 = 0.099 (solid). Note the non-monotonicity due to growth of f (p; γ) with diminution of γ counteracted by 1 the 1/γ power in equation (7g). Other parameters used: (p, q, m, s) = (3, 2, 2, 0), t× = 10, ǫ = 10 , x0 (0) = − 21 . Qualitatively similar results obtained with (p, q, m, s) = (2, 1, 2, 0), (2, 1, 3, 0), (4, 2, 2, 0). 261 With a single spike, or more generally any constellation with only one drift equation (more examples to be discussed hereinafter), the anomaly dependent acceleration or delay   (0) th plane. The slow do not affect the trajectory traversed by the i spike in the xi , Hi time variable τ parametrises this trajectory, hence transitions between different values of γ impact solely the velocity of the spike motion, as visualised in figure 5. In light of the above a general conclusion follows that for this degenerate pattern additional transitions will not result in worthy of investigation phenomena. 4.2. Pair of spikes A well explored constellation is the symmetric pattern, where the premise of reflective sym(0) (0) metry x0 < 0, x1 (τ ) ≡ −x0 (τ ) entails equal heights H0 (τ ) ≡ H1 (τ ) (Iron et al., 2001; Nec and Ward, 2012; Nec, 2016a). System (7) effectively reduces to one forthwith soluble 262 263 264 265 266 267 268 269 270 271 272 12 298 non-linear equation and one ordinary differential equation, whose closed form solution for γ = 1 is very similar to (8) and omitted here. Illatively from the single spike case, this pattern is not interesting as regards transient phenomena due to variable anomaly indices, as the trajectories are invariant albeit the parameterisation differs. Therefore it behoves one to centre the attention on asymmetric arrangements, whereat system (7) comprises two pairs of coupled equations. The effect of varying anomaly index is better evident when initially the loci are positioned significantly away from the symmetric constellation. Figure 6 shows the trajectories of (0) Hi (xi ), i = {0, 1} for a scenario of five distinct anomaly indices. Note the non-monotonicity of the left spike’s height adjustment: with diminishing γ in the first two transitions the spike grows higher, then lower as γ decreases further at the third transition, followed by slight growth with γ returning to normal. The response of the right spike does not manifest obvious conforming evolution, as it is located farther from its equilibrium point, its motion centred on attainment thereof. In particular the left spike overshoots the equilibrium height for a spell to compensate for the right spike’s position. Indeed eventually the two spikes are similarly close to the respective equilibria (compare locations at trajectory ends), albeit never symmetric. Juxtaposition with the propagation without anomaly reveals a twofold impact: the left spike, initially situated near its equilibrium point, would have achieved the equilibrium height very slowly but for the transition, whilst the right spike would have been nowhere close to the equilibrium position or height within the same time frame. Thus the transition to anomaly might be a powerful mechanism to speed up the pattern’s equilibration. With a set of kinetic exponents corresponding to higher values of β, and thence a stronger non-linearity, it is also possible to obtain a trajectory that includes an excursion away from equilibrium before converging thereto. This occurs when the initial heights of the two spikes are on different sides of the equilibrium value. Nonetheless here too the equilibration is essentially faster with anomaly. Figure 7 depicts a situation of this ilk. 299 4.3. Triple of spikes 273 274 275 276 277 278 279 280 281 282 283 284 285 286 287 288 289 290 291 292 293 294 295 296 297 300 301 302 303 304 305 306 307 308 309 310 311 312     The symmetric arrangement x0 (τ ), x1 (τ ), x2 (τ ) ≡ x0 (τ ), 0, −x0 (τ ) again effectively reduces system (7) equation of the autonomous type, to wit no differences   to a single differential (0) in trajectories xi (τ ), Hi (τ ) will ensue regardless of the parameterisation in τ . Thus, as before, only asymmetric constellations are of interest. Figure 8 gives a typical equilibration trajectory. Interesting phenomena to observe are earlier approach towards the equilibrium (0) height Heq for all three spikes with a transition to sub-diffusion, possible overshooting of (0) Heq whilst drift with regular diffusion begets no such occurrence (right spike x2 ), and fur(0) ther excursion from Heq by comparison to the regular trajectory (central spike x1 ). Figure 9 depicts the evolution for a combination of kinetic parameters that incur initial height on (0) different sides of the equilibrium value Heq . With this set of parameters, corresponding to a higher value of β, it is possible to obtain an overshooting of equilibrium by the normal trajectory, but not the anomalous one (left spike, x0 ), as well as a considerably delayed overshooting (right spike, x2 ). 13 4.4. Quadruple of spikes For any n > 3 a symmetric arrangement of either even   x0 (τ ), x1 (τ ), . . . xn/2−1 (τ ), −xn/2−1 (τ ) . . . , −x1 − x0 313 or odd   x0 (τ ), x1 (τ ), . . . x(n−1)/2−1 (τ ), 0, −x(n−1)/2−1 (τ ) . . . , −x1 − x0   number of spikes system (7) will require the solution of n2 non-linear equations for the spike heights and an equal number of respective drift differential equations. In this sense   (0) a pattern of four spikes is the first case, where a non-smooth trajectory xi (τ ), Hi (τ ) is expected with a symmetric initial state. Figure 10 shows such an example. With an asymmetric initial state similar phenomena are obtained as with n ⩽ 3: considerable excursions, faster occurring changes in spike height and marked non-monotonicity. Rather consistently the spikes drift a larger distance than subject to an uninterrupted normal diffusion regime. A typical example is given in figure 11. In light of these observations constellations of n > 4 are expected to introduce more variability within the available degrees of freedom, but reveal no qualitatively new phenomena. It is noteworthy that the same set of kinetic exponents that for pairs and triples engendered less typical evolution, with n = 4 failed to yield an equilibrated pattern (with initial conditions and all transition parameters identical to other sets used). This occurrence was further investigated and is discussed hereinafter. Figure 12 illustrates the smoothness of phase plane trajectories for the same constellation and initial conditions, but a discretised transition sequence approximating a smooth crossover in figure 1. From the result it is seen that a full solution would not produce tangibly smoother curves, proving the soundness of the approximation procedure. The computation was performed for all foregoing constellations with a similar aftermath. The particular example shown was chosen due to well defined non-monotonicity in height evolution that was nonetheless traced smoothly. 5. Existence of solutions 314 315 316 317 318 319 320 321 322 323 324 325 326 327 328 329 330 331 332 333 334 335 (0) System (7) is valid as long as Hi > 0 ∀i and all xi are sufficiently far apart. Being a reduction of the full partial differential system (2) on the slow time scale τ , (7) cannot track behaviour that breaches the assumptions of quasi-equilibrium. One such event of relevance here is a termination of a solution branch due to a bifurcation. To the best of the author’s knowledge this occurrence has not been formally explored in the algebraic-differential system associated with the spike solution in Gierer-Meinhardt model, albeit it has been observed in numerical (Tzou, 2016). A constellation following   simulations (0) a branch in the 2n-dimensional space xi , Hi , i = {0, . . . , n − 1} that terminates before the equilibrium point is reached, will not be able to equilibrate. Only simulation of the full partial differential system (2) can elucidate the shift to another solution branch and 336 337 338 339 340 341 342 343 344 345 14 370 possible changes in the constellation, a study beyond the ambit of the current analysis. The existence and concomitant termination of the trajectories in each of   of a bifurcation (0) is an inherent property of the reduced Gierer-Meinhardt system and the planes xi , Hi not the result of sub-diffusion or transitions between distinct values of the anomaly index, however a bifurcation might induce an earlier termination in the presence of anomaly or beget unbounded growth. Typical examples appear hereunder, and no attempt is made to map the parameter space effecting a bifurcation. The first example is for a system with regular diffusion and no transitions, i.e. γ ≡ 1. (0) Figure 13 shows the spike velocity x′i (τ ) and trajectories Hi (xi ) for a quadruple of spikes. Whilst the velocity is finite, note that the slope at the terminus point τtrm becomes infinite, as d (0) does H . Nevertheless x′i (τ ) remains finite, as this is not a singularity of (7g). One must dτ i bear in mind the terminus is indeed attained – the integration ceases because (7a) possesses no solution, in contrast to the equilibration itself with its exponentially or algebraically slow approach towards the fixed point, or a finite time blow up, where computation becomes impractical. If γ is moderately small, a constellation might reach a branch terminus, whilst with γ ≡ 1 no bifurcation is encountered. An example of this is given in figure 14. This constellation is initialised identically to the one used in figure 11 with sets of kinetic exponents begetting no solution existence issues. An unimpeded growth of x′i (τ ) might ensue for small values of γ, when the right hand γ side of (7g) exceeds unity in magnitude. Re-write the left hand side of (7g) as sgn x′i x′i , whence it is readily inferred that 1/γ begets a power growth. For instance, with the same set of parameters as in figure 14, if another cross-over moment is added anywhere in the range t× 1 < t× 2 < ttrm and γ3 < γ2 , the solution grows beyond the computer’s handling ability within a few integration steps after the transition. 371 6. Discussion 346 347 348 349 350 351 352 353 354 355 356 357 358 359 360 361 362 363 364 365 366 367 368 369 372 373 374 375 376 377 378 379 380 381 The infinitesimally narrow spike solution, existing for the Gierer-Meinhardt model with both regular diffusion (γ = 1) and sub-diffusion (0 < γ < 1), persists when the index γ, traditionally a constant, is a piecewise constant function. To wit, the validity of the asymptotic spike pattern extends to the situation when the system undergoes transitions between regimes, whose mean square displacement evolution hr2 (t)i ∼ tγ is characterised by distinct exponents. The number of transition points as well as their cross-over timings t× j and corresponding anomaly indices 0 < γj ⩽ 1 are arbitrary. At each transition γj −→ γj+1 occurring at t× j , j = 1, 2, . . ., the drift velocity x′i (τ ) has a (0) jump discontinuity, therefore the spike height Hi (τ ) and locus position xi (τ )are continuous,  (0) but not smooth. Generally the non-smoothness is retained in the plane xi , Hi , with (0) 382 383 the exception of trajectories Hi (xi ) where τ is an arbitrary parameterisation variable, i.e. for those constellations that are effectively described by a sole non-linear equation for the 15 height and respective differential equation for locus drift, namely a single spike, a pair of symmetric spikes, and a triple with central spike at equilibrium x1 ≡ 0 and outer spikes moving symmetrically. Thence for n ⩽ 3 only asymmetric initial states yield qualitatively distinct trajectories for a sequence of transitions by comparison to a constant γ(t). For n ⩾ 4 both symmetric and asymmetric initial states give non-smooth trajectories in the locus position – spike height planes. Near the equilibrium points transitions occasion but small differences in velocity, therefore often the non-smoothness is virtually indiscernible. The manifold alterations in the trajectory form range from minor and quantitative to qualitatively novel behaviour. Depending on the specific parameters chosen, a pattern might equilibrate significantly faster than with a single fixed anomaly index, be it normal or subdiffusive; the spikes might drift a larger distance in the process; the spike height might reach a local maximum or minimum that would not have been attained without a series of transitions. Recollecting that the spike height conforms to reactant concentration in the original chemical model, these results bear on recent experimentally documented concepts related to reactions within such complex biological media as cell membranes, organelles and nuclei. The most prominent aftermath of permitting the conventionally constant anomaly index to vary in time is the instantaneous adjustment of the spike’s velocity. This behaviour qualitatively corresponds to a number of scenarios. One is a concatenation of events, when a cascade of consecutive periods of molecule diffusion and subsequent reactions trigger each other in turn, each with its own physical characteristics due to the locality of each drifting cluster and ligands. When out of a set of identical diffusing molecules a sufficient number is bound at the target, the concentration of that reactant diminishes, signalling the next reaction in the chain and possible slowing down of the initial diffusion. Another is that docking of the designated molecules will trigger a fast removal of the remaining reactant, thereby requiring a significant acceleration of their spatial propagation. Yet another is diffusion through distinct media, like penetration in and out of a nucleus or a cell or within such systems as the lymphatic vessels and nodes, where multifarious stages can be characterised by different anomaly indices. The findings presented are based on an asymptotic reduction of a system of partial differential equations of a fractional order. Ideally it is desirable to obtain a numerical verification in those parts of the parameter space, where the asymptotic solution is valid. Numerical solution of partial fractional differential equations is notoriously difficult and often defies conventional concepts of numerical analysis as regards consistence, accuracy and stability of schemes (Meerschaert and Tadjeran, 2004). Here the challenge due to the kernel singularity of (2e) is further compounded by an increase in the spike width with diminishing γ, whereby the interval of feasible values of ǫ is narrower than the concomitant expected for normal diffusion and integer order equations. Development of a suitable numerical scheme is an enduring problem and topic of future research. All past successful schemes resorted to a regularisation of the fractional operator, a course of action impossible here, given that the operator is needed in its current form for the spike pattern to exist as well as certain spectral properties to be retained. The Gierer-Meinhardt system is a paradigmatic model, representing no particular natural application, but capable of reconstructing general observable phenomena. Whist having once 384 385 386 387 388 389 390 391 392 393 394 395 396 397 398 399 400 401 402 403 404 405 406 407 408 409 410 411 412 413 414 415 416 417 418 419 420 421 422 423 424 425 16 448 more proved a powerful pattern forming tool, the limitation of the solution obtained here is in the requirement that the anomaly index remain constant between any two transition points. Whilst a formally continuous variation of γ might be a curious topic for future research, the posed theory was shown to yield evolution trajectories appearing visibly smooth when a smooth function γ(t) was discretised into a sequence of step-like segments. The limitations on the validity of the presented asymptotic solution are not burdensome, therefore such discretisation is feasible for most reasonably behaved γ(t). Moreover, the method itself is generic and might be easily adapted to unrelated dynamical systems, as long as an asymptotic solution with fixed scales is available. The degree of smoothness in the solution will depend on the particular system at hand. It is furthermore conjectured that for γ(t) of an abrupt, yet continuous variation, for instance an inner layer as in (6) with ε ≪ 1, away from the transition core the solution obtained herein will faithfully capture the dynamics, whilst in the vicinity thereof some delay in adjustment of the spike velocity will ensue, depending on the specific dependence of γ(t). Perhaps an asymptotic solution might be furnished for certain forms of γ(t) based on an independent small parameter describing the abruptness of transition. With an unrestricted variation, i.e. γ(t) such that no regions can be approximated by a constant, a disparate method of solution will be required, as the current asymptotic approach becomes unfeasible. The Gierer-Meinhardt system is further amenable to a generalisation involving a fractional Laplacian (spatial fractional derivative), thereby admitting solutions of spikes that propagate according to Lévy flights, a type of super-diffusion (Nec, 2012). Extension of the current analysis to include transitions between Lévy flights with distinct anomaly indices as well as between sub-diffusive to super-diffusive regimes is another future problem. 449 References 426 427 428 429 430 431 432 433 434 435 436 437 438 439 440 441 442 443 444 445 446 447 450 451 452 453 454 455 456 457 458 459 Bakalis, E., Höfinger, S., Venturini, A., 2015. Crossover of two power laws in the anomalous diffusion of a two lipid membrane. J. Chem. Phys. 142, 215102. Cabal, G. G., Genovesio, A., Rodrigues-Navarro, S., Zimmer, C., Gadal, O., Lesne, A., Buc, H., Feuerbach-Fournier, F., Olivo-Marin, J. C., Hurt, E. C., Nehrbass, U., 2006. SAGA interacting factors confine sub-diffusion of transcribed genes to nuclear envelope. Nature 441, 770–773. Chen, W., Sun, H., Zhang, X., Korošak, D., 2010. Anomalous diffusion modeling by fractal and fractional derivatives. Comp. Math. Appl. 59, 1754–1758. Chen, W., Zhang, J., Zhang, J., 2013. A variable-order time-fractional derivative model for chloride ions sub-diffusion in concrete structures. Frac. Calc. Appl. Analysis 16, 76–92. 461 Codling, E. A., Plank, M. J., Benhamou, S., 2008. Random walk models in biology. J. R. Soc. Interface 5, 813–834. 462 Eliazar, I., Klafter, J., 2011. Anomalous is ubiquitous. Annals of Physics 326, 2517–2531. 460 17 Elliot, D., 1993. An asymptotic analysis of two algorithms for certain Hadamard finite-part integrals. IAM J. Num. Analysis 13, 445–462. Gierer, A., Meinhardt, H., 1972. A theory of biological pattern formation. Kybernetik 12, 30–39. Henry, B. I., Wearne, S. L., 2000. Fractional reaction–diffusion. Physica A 276, 448–455. Iron, D., Ward, M. J., Wei, J., 2001. The stability of spike solutions to the one-dimensional Gierer-Meinhardt model. Physica D 150, 25–62. Klages, R., Radons, G., Sokolov, I. M., 2008. Anomalous transport: foundations and applications. Wiley-VCH, Weinheim. McKinley, S. A., Yao, L., Forest, M. G., 2009. Transient anomalous diffusion of tracer particles in soft matter. J. Rheol. 53, 1487. Meerschaert, M. M., Nane, E., Vellaisamy, P., 2013. Transient anomalous sub-diffusion on bounded domains. Proc. Amer. Math. Soc. 141, 699–710. Meerschaert, M. M., Tadjeran, C., 2004. Finite difference approximations for fractional advection – dispersion flow equations. J. Comput. Appl. Math. 172, 65–77. Metzler, R., Klafter, J., 2000. The random walk’s guide to anomalous diffusion: a fractional dynamics approach. Physics Reports 339, 1–77. Morgan, N. A., Spera, F. J., 2001. A molecular dynamics study of the glass transition in CaAl2 Si2 O8 : thermodynamics and tracer diffusion. American Mineralogist 86, 915–926. Naber, M., 2004. Distributed order fractional sub-diffusion. Fractals 12, 23. Nec, Y., 2012. Spike-type solutions to one dimensional Gierer-Meinhardt model with Lévy flights. Studies in Appl. Math. 129, 272–299. Nec, Y., 2016a. Dynamics of spike solutions in Gierer-Meinhardt model with time dependent diffusivity. preprint. Nec, Y., 2016b. Explicitly solvable eigenvalue problem and bifurcation delay in sub-diffusive Gierer-Meinhardtmodel. Euro. J. Appl. Math., doi:10.1017/S0956792516000012. Nec, Y., Ward, M. J., 2012. Dynamics and stability of spike-type solutions to a one dimensional Gierer-Meinhardt model with sub-diffusion. Physica D 241, 947–963. 463 464 465 466 467 468 469 470 471 472 473 474 475 476 477 478 479 480 481 482 483 484 485 486 487 488 489 490 Oldham, K. B., Spanier, J., 1974. Fractional Calculus. Academic Press, New York. 491 Ramirez, L. E. S., Coimbra, C. F. M., 2009. On the selection and meaning of variable order operators for dynamic modeling. International J. Diff. Eq. 2010, 846107. 493 492 18 494 495 496 497 498 499 500 501 502 503 504 Reynolds, A. M., 2005. On the anomalous diffusion characteristics of membrane-bound proteins. Physics Letters A 342, 439–443. Saxton, M. J., 1997. Single-particle tracking: the distribution of diffusion coefficients. Biophys. J. 72, 1744–1753. Saxton, M. J., 2007. A biological interpretaton of transient anomalous subdiffusion. i. qualitative model. Biophys. J. 92, 1178–1191. Sun, H., Chen, W., Chen, Y. Q., 2009. Variable-order fractional differential operators in anomalous diffusion modeling. Physica A 388, 4586–4592. Sun, W., Ward, M. J., Russell, R., 2005. The slow dynamics of two-spike solutions for the Gray-Scott and Gierer-Meinhardt systems: competition and oscillatory instabilities. SIAM J. Appl. Dyn. Sys. 4(4), 904–953. 507 Torreno-Pina, J. A., Manzo, C., Garcia-Parajo, M. F., 2016. Uncovering homo- and heterointeractions on the cell membrane using single particle tracking approaches. J. Phys. D: Appl. Phys. 49, 104002. 508 Tzou, J. C., 2016. private communication. 505 506 509 510 511 Tzou, J. C., Bayliss, A., Matkowsky, B. J., Volpert, V. A., 2011. Stationary and slowly moving localised pulses in a singularly perturbed Brusselator model. Euro. J. Appl. Math. 22(5), 423–453. (0) x0 (τ ) H0 (τ ) 19 (0) H0 (τ× ) x0 (τ× ) (0) H0 (0) x0 (0) τ× τ× 0 τ τ H0 (0) 0 (0) H0 (τ× ) (0) H0 (0) x0 (0) x0 (τ× ) x0 Figure 5: Trajectory parameterisation example for a single spike and transition scenario γ(t < t× ) = 1, γ(t > t× ) = 15 (0) (solid curves) and drift with no transition, γ ≡ 1 (dotted curves): locus position x0 (τ ) (upper left panel), H0 (τ ) (upper (0) right panel) and H0 (x0 ) (bottom panel). Diamonds mark the cross-over moment τ× = t× ǫγ(t)+1 on all panels. Circle marks the end of the indistinguishable dotted curve on the bottom panel: from the upper panels it is obvious that the spike speeds up upon transition, therefore the solid trajectory (with transition) extends further than the dotted one (no transition). 1 , x0 (0) = − 43 . Qualitatively similar results obtained with Other parameters used: (p, q, m, s) = (3, 2, 2, 0), t× = 50, ǫ = 10 (p, q, m, s) = (2, 1, 2, 0), (2, 1, 3, 0), (4, 2, 2, 0). 20 (0) H0 (0) Heq (0) H0 (0) -0.42 x0 x0 (0) (0) H1 (0) Heq (0) H1 (0) x1 (0) 0.42 x1 Figure of five intervals characterised by  o parameterisation example for a pair of spikes and transition  scenario   n 6: Trajectory 3 1 1 . Equilibrium position , 10 , 1 with cross-over moments t× = {1, 2, 6, 8}. Initial position is x0 , x1 = − 52 , 10 γ = 1, 41 , 20     is x0 , x1 = − 21 , 12 . Equilibrium height, equal for both spikes, is marked by the dashed line. Dotted lines show the trajectory for γ ≡ 1 (in upper panel continuing beyond the frame given). Diamonds mark the cross-over moments t× j . Other 1 parameters used: (p, q, m, s) = (3, 2, 2, 0), ǫ = 10 . Qualitatively similar results obtained with (p, q, m, s) = (2, 1, 2, 0), (4, 2, 2, 0). 21 (0) H0 (0) max H0 (0) Heq x0 (t× 4 ) x0 x0 (0) (0) H0 (0) max H0 (0) H0 (0) x0 (t× 3 ) x0 (t× 2 ) x0 x0 (t× 1 ) 22 (0) H1 Heq (0) H1 (0) x1 (t× 1 ) x1 (t× 4 ) x1 Figure of non-monotonic equilibration for a pair of spikes with transition scenario of five intervals characterised by n 7: Example o 3 1 , 10 , 1 with cross-over moments t× = {1, 2, 6, 8}: left spike exhibits trajectory excursion away from equilibrium γ = 1, 41 , 20 prior to convergence (upper enlarged in central  panel), right spike’s  response is monotonic (lower   panel, height  peak vicinity panel). Initial position is x0 , x1 = 1 . Equilibrium position is − 52 , 10 x0 , x1 = − 12 , 21 . Equilibrium height, equal for both spikes, is marked by the dashed line (observe left spike begins above equilibrium height, whilst right spike is below). Dotted lines show the trajectory for γ ≡ 1 (in upper and central panels continuing beyond the frame given). Diamonds mark 1 the cross-over moments t× j . Other parameters used: (p, q, m, s) = (2, 1, 3, 0), ǫ = 10 . 23 (0) H0 (0) Heq (0) H0 (t× 1 ) x0 (t× 4 ) x0 x0 (t× 1 ) (0) H1 (0) max H1 (0) Heq (0) H1 (t× 1 ) x1 (t× 4 ) x1 x1 (t× 1 ) 24 (0) H2 (0) max H2 (0) Heq (0) H2 (t× 1 ) x2 (t× 1 ) x2 x2 (t× 4 ) Figure by  o parameterisation example for a triple of spikes and transition scenario of  five  intervals characterised n 8: Trajectory 3 1 1 3 . Equilibrium , 10 , 1 with cross-over moments t× = {1, 2, 6, 25}. Initial position is x0 , x1 , x2 = − 35 , 10 , 10 γ = 1, 41 , 20     position is x0 , x1 , x2 = − 32 , 0, 23 . Equilibrium height, equal for all spikes, is marked by the dashed line. Dotted lines show 1 the trajectory for γ ≡ 1. Diamonds mark the cross-over moments t× j . Other parameters used: (p, q, m, s) = (3, 2, 2, 0), ǫ = 10 . Qualitatively similar results obtained with (p, q, m, s) = (2, 1, 2, 0), (4, 2, 2, 0). 25 (0) H0 (0) Heq (0) H0 (t× 1 ) x0 x0 (t× 4 ) x0 (t× 1 ) (0) H1 (0) max H1 (0) Heq x1 (t× 4 ) x1 x1 (t× 1 ) 26 (0) H2 (0) max H2 (0) Heq (0) H2 (t× 1 ) x2 (t× 1 ) x2 x2 (t× 4 ) Figure by n 9: Trajectory  o parameterisation example for a triple of spikes and transition scenario of  five  intervals characterised 3 1 1 3 γ = 1, 41 , 20 . Equilibrium , 10 , 1 with cross-over moments t× = {1, 2, 6, 25}. Initial position is x0 , x1 , x2 = − 35 , 10 , 10     position is x0 , x1 , x2 = − 32 , 0, 23 . Equilibrium height, equal for all spikes, is marked by the dashed line. Dotted lines show 1 the trajectory for γ ≡ 1. Diamonds mark the cross-over moments t× j . Other parameters used: (p, q, m, s) = (2, 1, 3, 0), ǫ = 10 . 27 (0) max H0 (0) H0 (0) Heq (0) H0 (0) x0 (t× 1 ) x0 x0 (t× 4 ) (0) H1 (0) Heq (0) H1 (t× 1 ) x1 (t× 1 ) x1 x1 (t× 4 ) Figure 10: Trajectoryn parameterisation scenario  o example for a symmetric quadruple of spikes and transition   of five intervals 3 1 19 , 10 , 1 with cross-over moments t× = {8, 10, 20, 30}. Initial position is x0 , x1 = − 20 , − 34 . characterised by γ = 1, 41 , 20     (0) (0) Equilibrium position is x0 , x1 = − 43 , − 14 . The two rightmost spikes satisfy x2 (τ ) ≡ −x1 (τ ), H2 (τ ) ≡ H1 (τ ), (0) (0) x3 (τ ) ≡ −x0 (τ ), H3 (τ ) ≡ H0 (τ ). Equilibrium height, equal for all spikes, is marked by the dashed line. Dotted lines show 1 the trajectory for γ ≡ 1. Diamonds mark the cross-over moments t× j . Other parameters used: (p, q, m, s) = (3, 2, 2, 0), ǫ = 10 . Qualitatively similar results obtained with (p, q, m, s) = (2, 1, 2, 0), (2, 1, 3, 0), (4, 2, 2, 0). 28 (0) max H0 (0) H0 (0) Heq (0) H0 (t× 1 ) x0 (t× 1 ) x0 x0 (t× 4 ) (0) H1 (0) Heq (0) H1 (t× 1 ) x1 (t× 1 ) x1 x1 (t× 4 ) 29 (0) H2 (0) Heq (0) min H2 x2 (t× 4 ) x2 x2 (t× 1 ) (0) H3 (0) H3 (t× 1 ) (0) min H3 x3 (t× 4 ) x3 x3 (t× 1 ) Figure 11: example for a quadruple of spikes and transitionscenario of five characterised  o  intervals  n Trajectory parameterisation 3 1 19 9 . , 10 , 1 with cross-over moments t× = {1, 2, 6, 8}. Initial position is x0 , x1 , x2 , x3 = − 20 , − 12 , 53 , 10 by γ = 1, 41 , 20     3 1 1 3 Equilibrium position is x0 , x1 , x2 , x3 = − 4 , − 4 , 4 , 4 . Equilibrium height, equal for all spikes, is marked by the dashed line. Dotted lines show the trajectory for γ ≡ 1. Diamonds mark the cross-over moments t× j . Other parameters used: 1 (p, q, m, s) = (3, 2, 2, 0), ǫ = 10 . Qualitatively similar results obtained with (p, q, m, s) = (2, 1, 2, 0), (4, 2, 2, 0). With the set (p, q, m, s) = (2, 1, 3, 0) the pattern failed to equilibrate. H0 (0) 30 (0) Heq (0) H0 (0) 19 − 20 −0.82 x0 (0) H1 (0) Heq (0) H1 (0) − 21 −0.32 x1 31 (0) H2 (0) H2 (0) (0) min H2 0.32 x2 3 5 (0) H3 (0) H3 (0) (0) Heq 0.82 x3 9 10 Figure 12: Trajectory parameterisation example for a quadruple of spikes and inner layer transition scenario (figure 1) with     3 3 9 γ0 = , γ1 = and ε = 5 in (6). Initial position is x0 , x1 , x2 , x3 = − 19 . Equilibrium position is , − 12 , 35 , 10 20 4    20 x0 , x1 , x2 , x3 = − 34 , − 41 , 14 , 43 . Diamonds mark the cross-over moments t× j . Other parameters used: (p, q, m, s) = 1 (3, 2, 2, 0), ǫ = 10 . Qualitatively similar results obtained with (p, q, m, s) = (2, 1, 2, 0), (4, 2, 2, 0). 32 x′i (τ ) 1 0 -1 τtrm 0 τ (0) Hi (0) max Hi (0) min Hi -1 0 xi Figure 13: Example of trajectory termination due to a bifurcation for a quadruple of spikes with γ ≡ 1. Top panel: velocity x′i (τ ), i = 0 (leftmost spike, solid curve), i = 1 (left central spike, dashed curve), i = 2 (right central spike, dash-dotted (0) (0) curve) and i = 3 (rightmost spike, dotted curve). Bottom panel: Hi (xi ) for the left spikes, i = {0, 1}, and Hi (−xi ) for the  right ones, i ={2, 3}. Diamonds  mark trajectory termini.  Curve linestyles  are identical to top panel. Initial position is 9 − 19 . Equilibrium position x0 , x1 , x2 , x3 = , − 12 , 53 , 10 20 − 43 , − 41 , 14 , 43 is never reached, the terminus 1 ′ point τtrm characterised by infinite slope of xi (τ ). Other parameters used: (p, q, m, s) = (3, 3, 3, 0), ǫ = 10 . x0 , x1 , x2 , x3 = 33 x′0 (τtrm ) x′1 (τ ) x′0 (τ ) x′1 (τtrm ) x′0 (0) x′1 (0) τtrm τ× 1 τtrm τ× 1 τ τ x′3 (τtrm ) x′3 (τ ) x′2 (τ ) x′2 (τtrm ) x′2 (0) x′3 (0) τtrm τ× 1 τtrm τ× 1 τ τ (0) H0 (ttrm ) (0) (0) H1 H0 (0) H1 (ttrm ) (0) H0 (t× 1 ) (0) H1 (t× 1 ) x0 (t× 1 ) x1 x0 (ttrm ) (0) H2 (ttrm ) x1 (t× 1 ) x1 (ttrm ) x2 (0) (0) H3 H2 (0) H3 (ttrm ) (0) (0) H3 (t× 1 ) H2 (t× 1 ) x2 (t× 1 ) x3 x2 (ttrm ) x4 x3 (t× 1 ) Figure 14: Example of trajectory termination due to a bifurcation for a quadruple of spikes with γ = x3 (ttrm ) n 1, 14 o and cross-over moment t× = 1. Top: velocity x′i (τ ). Solid curves: i = 0 (leftmost spike), i = 1 (left central spike), i = 2 (right central spike) (0) and i = 3 (rightmost spike). Dotted curves show the same evolution for γ ≡ 1. Bottom: respective Hi (xi ). Diamonds mark     19 9 trajectory termini. Crosses mark the cross-over moment. Initial position is x0 , x1 , x2 , x3 = − 20 . Equilibrium , − 12 , 53 , 10     3 1 1 3 position x0 , x1 , x2 , x3 = − 4 , − 4 , 4 , 4 is never reached, the terminus point τtrm characterised by infinite slope of x′i (τ ) 1 . for the the left pair of spikes, x′i (τtrm ) = 0 for the right pair. Other parameters used: (p, q, m, s) = (2, 1, 3, 0), ǫ = 10