Theoretical Perspectives on Biological Machines

Mauro L. Mugnai∗ Department of Chemistry, University of Texas, Austin, TX 78712 Changbong Hyeon† Korea Institute for Advanced Study, Seoul 02455, Republic of Korea Michael Hinczewski‡ Department of Physics, Case Western Reserve University, OH 44106 D. Thirumalai§ Department of Chemistry, University of Texas, Austin, TX 78712

Many biological functions are executed by molecular machines, which like man made motors consume energy and convert it into mechanical work. Biological machines have evolved to transport cargo, facilitate folding of and RNA, remodel chromatin and replicate DNA. A common aspect of these machines is that their functions are driven by fuel provided by hydrolysis of ATP or GTP, thus driving them out of equilibrium. It is a challenge to provide a general framework for understanding the functions of bi- ological machines, such as molecular motors (, , and ), molecular chaperones, and . Using these machines, whose structures have little resem- blance to one another, as prototypical examples, we describe a few general theoretical methods that have provided insights into their functions. Although the theories rely on coarse-graining of these complex systems they have proven useful in not only accounting for many in vitro experiments but also address questions such as how the trade-off be- tween precision, energetic costs and optimal performances are balanced. However, many complexities associated with biological machines will require one to go beyond current theoretical methods. We point out that simple point mutations in the enzyme could drastically alter functions, making the motors bi-directional or result in unexpected dis- eases or dramatically restrict the capacity of molecular chaperones to help proteins fold. These examples are reminders that while the search for principles of generality in biology is intellectually stimulating, one also ought to keep in mind that molecular details must be accounted for to develop a deeper understanding of processes driven by biological machines. Going beyond generic descriptions of in vitro behavior to making genuine understanding of in vivo functions will likely remain a major challenge for some time to come. In this context, the combination of careful experiments and the use of physics and physical chemistry principles will be useful in elucidating the rules governing the workings of biological machines.

CONTENTS IV. Stochastic Kinetic Models 7 A. The Master Equation 8 I. Introduction 2 B. Thermodynamics 9 C. Rate of Entropy Production 10 arXiv:1908.11323v1 [physics.bio-ph] 29 Aug 2019 II. Structures 4 D. Fluctuation Theorems 10 A. Molecular motors 4 B. Catalytic Cycle 5 V. Molecular Motors – Models without Detachment 12 A. Molecular Motors are Processive Enzymes 12 III. Biological machines are active systems. 6 B. Processive Motor Velocity 13 C. Periodic Lattice Model 13 1. One-state Models 14 2. Multi-state, Uni-cycle Models 15 3. Multi-cycle Models 17 ∗ [email protected] D. Additional remark on non-equilibrium nature of † [email protected] motors 18 ‡ [email protected] E. Applications 19 § [email protected] 1. Myosin V and Kinesin-1 19 2

2. Dynein – an erratic motor 20

VI. Molecular Motors – Models with Detachment 20 A. Velocity Distribution 21 B. Alternate Models with Detachment 22

VII. Polymer Physics-based Approaches Incorporating Structural Features Into Kinetic Theories 23

VIII. Simulations using Coarse-Grained Models 28

IX. Cost-precision Trade-off and Efficiency of Molecular Motors 30 1. Cost-precision trade-off and its physical bound of molecular motors 30 2. Transport Efficiency 31

X. Molecular chaperones 33

XI. Universal characteristics of Helicases 36 FIG. 1 Illustration of the complexity of transportation of , which are vesicles containing melanin. Both !1 XII. Discussion 38 dynein and kinesin-2 compete for the same binding site on A. Sometimes Details Matter 38 mediated by p150Glued. Depending on the func- B. Efficiency and optimality 38 tion (aggregation of melanosomes or their dispersion through- C. Specificity versus Promiscuity 39 out the cell) one or the other wins. When melanosomes are D. Biological complexity 39 dispersed in the cell they are transported by kinesin-2 and XIII. A Final remark 39 myosin V whereas when they aggregate dynein moves the cargo. Figure extracted from Soldati and Schliwa (2006). Acknowledgments 40

References 40 and abiotic systems. Because of functional demands in living systems, which also includes adaptation to chang- I. INTRODUCTION ing environmental conditions, it is virtually impossible to fully describe biology without evolutionary considera- The opening sentence of a perspective by Bustamante tions. Although not the focus of our perspective, the role (Bustamante et al., 2011) on the workings of nucleic acid of evolutionary constraints must also be integrated with translocases begins with the quote, “The operative indus- physical models in order to discover general principles try of Nature is so prolific that machines will be eventu- governing the functions of biological machines. ally found not only unknown to us but also unimaginable What are the characteristics of biological machines? by our mind”, attributed to Marcello Malpighi, who is First, there are many varieties of machines, all of which considered the founder of microscopic anatomy, histol- should be thought of as enzymes, which consume energy ogy, and embryology. This statement, made over three and perform mechanical work to carry out specific tasks. centuries ago, is even more relevant today. It is a re- A few of the machines that we consider here are kinesin, minder that mechanical forces must play a fundamental myosin, and dynein, which are collectively referred to as role in biology. In modern times this vast subject falls molecular motors (Belyy et al., 2014a; Block, 2007; Reck- under the growing field of mechanobiology. The perva- Peterson et al., 2018; Sun and Goldman, 2011a; Vale sive role of mechanics in living systems controls motil- and Milligan, 2000). These cytoskeletal motors trans- ity on all length scales, from of a single cell on port cargo by walking, almost always unidirectionally, a substrate and collections of cells to dynamics at the on filamentous and (MT). Under in molecular level. Cooperative interactions between var- vivo conditions the motors cooperate or there could be a ious modules at the molecular level is thought to con- tug-of-war in the process of transport (see Fig. 1) (Gross trol functions at the mesoscale. A number of complex et al., 2002; Levi et al., 2006). To illustrate the diversity dynamical processes such as transcription, translation, of motor-like functions and point out certain emerging transport of vesicles and , folding of proteins general principles, we also provide theoretical descrip- and RNA, and chromosome segregation control the sus- tions for the functions of molecular chaperones, which tenance and growth of cells. At some level all these bio- assist in the folding of proteins and ribozymes (RNA en- logically important processes involve molecular machines, zymes) that cannot do so spontaneously, and helicases whose ability to perform their functions, often but always with multiple functions that includes separation of dou- with high efficiency, in a noisy crowded environment is ble stranded (ds) DNA strands. Both molecular chap- truly remarkable. It is worth remembering that the abil- erones and helicases bear no structural resemblance or ity to execute a variety of functions distinguishes living sequence similarity to molecular motors. Nevertheless, 3 we are convinced that by comparing the characteristics of these seemingly unrelated machines, integrating the- ory and experiments, unifying themes and differences be- tween them could be elucidated. Second, all these motors and others not covered here (for example polymerases, , and packaging motors) are all multidomain proteins, whose architectures are spectacularly different. Despite the vastly different sequences, structures, and evolutionary origins it is indeed the case that all of these machines utilize some form of chemical energy (gener- ated by ATP or GTP hydrolysis) in order to amplify the small local conformational changes through their struc- tural linkages to facilitate large conformational changes for functional purposes. Such conformational amplifica- tions are examples of remarkable allostery or action at a distance, and could also be couched in terms of in- FIG. 2 Significant physical timescales associated with myosin V dynamics, from the coarse-grained polymer theory model formation transfer between structural subunits that are of Hinczewski et al. (2013). spatially well separated, which in some cases (dynein for example (Bhabha et al., 2014)) can be as large as 25nm. A large number of experiments have unveiled many of the details of how these machines move by converting ular dynamics simulations have been performed to get a chemical energy into mechanical work (see for example molecular picture of certain aspects of the functions of (Hartman et al., 2011; Spudich and Sivaramakrishnan, motors (Hwang et al., 2008, 2017) and other machines 2010) (De La Cruz and Ostap, 2004; Holzbaur and Gold- (see for example (Elber and West, 2010; Ma et al., 2000; man, 2010; Sweeney and Houdusse, 2010). A remarkable Stan et al., 2005), the current limitations of such ap- number of experimental methods have been developed proaches prevent them from making direct contact with to address various aspects of biological machines. These experiments. Improvements in the development of ac- include, but are not restricted to, ensemble experiments curate energy functions (also known as force fields) and (stopped flow and fluorescent labeling) that provide the enhancement in computer capacity to enable simulations much needed data on ATP hydrolysis and ADP release for long times will in the future bridge the gap between rates, single molecule experiments that yield dwell time what is currently possible and what is needed for realistic distributions in molecular motors, processivity and veloc- description of molecular machines. ity as a function of external loads in motors and helicases. The actions of molecular machines, like all biological In addition, a combination of ensemble experiments and processes, are complicated, involving cooperative dynam- determination of structures using X-ray crystallography ics on multiple time scales. This is illustrated using the and cryo-EM experiments has produced a vivid picture of typical time scales involved in a single step of myosin V the molecular basis of chaperone function. Of particular (Fig. 2). In many problems in physics it is the hoped that note are optical trap experiments, which have been used simple models, that capture the essence of the problem to obtain mean motor velocity as a function of a resis- can be created, which be can be solved, in order to ob- tive force and ATP concentration in kinesin, myosin, and tain insights into highly non-trivial systems. This strat- dynein, and the dependence of velocity and processivity egy is hard to implement for biological machines because in a number of helicases. No point would be served in the interaction energies are highly heterogeneous (like in reviewing the detailed results from these experiments as spin glasses), thus making it hard to construct a reason- there are many articles that the interested reader might able coarse-graining strategy. Nevertheless, in order to consult. make progress one has to devise tractable coarse-grained Our focus here is to describe theoretical approaches models, which can be either simulated or solved ana- that are rooted in a number of areas in physics in order to lytically (at least approximately). The efficacy of such understand principally outcomes of in vitro experiments. approaches can then be assessed by direct comparisons The theoretical approaches in these studies were devel- with experimental results, and their abilities to provide oped, and continue to be the focus of current research, insights into the mechanisms of their functions. Remark- in order to quantitatively explain the experimental ob- ably, inspired by advances in experiments in the last servations, and shed light on new puzzles that seem to decade, several theoretical models have been proposed, arise with ever improving advances in experimental tech- which have greatly contributed to our understanding of niques. Perhaps, the most detailed view of how biological molecular machines. These developments have occurred machines operate might be obtained from molecular dy- in different contexts, which belie the underlying unifying namics simulations. Although atomically detailed molec- principles. To bring these issues to the fore, we pro- 4 vide our perspectives on the application of these models, which should be thought of as coarse-grained network a b c d GroEL-GroES models or their generalizations. Such models have been Kinesin Dynein created for explaining not only the motility of motors, 23 nm but also the functions of molecular chaperones, and he- Myosin V licases. We focus on the applications of these theoretical ideas in order to account for the functions of these in- 14 nm trinsically non-equilibrium systems. Collectively, these approaches show that, by examining in detail the work- e ings of many machines, universal principles, both at the 8 nm 20 nm conceptual and practical levels, might emerge.

The literature on the functions of biological machines 10 nm is vast. Therefore, we restricted ourselves to only a few 35 nm topics that focus on theoretical approaches, which are FIG. 3 Schematic representations of the structures of six bio- sufficiently general that they can be applied to an ar- logical machines. Molecular motors, like conventional kinesin ray of problems in the field. Rather than describe many (a), dynein (b), and myosin V (c), show great variations in results in detail, we walk the the reader through a se- their structures and sizes. The motor heads in kinesin and lection of theoretical methods, which is necessarily bi- myosin V, shown in dark blue, bind directly to the micro- ased, so that she or he can access the literature read- tubule and actin, respectively (not shown) but in dynein the binding domain (light blue) is separated from the ily. Before getting to the details of this review we will motor domain (hexameric ring structures) by nearly twenty be remiss if we did not point out one prescient mono- five nanometers. (d) Structure of the GroEL in the symmet- graph (Howard, 2001) and two forward looking reviews ric state ( Data Bank (PDB) code 4PKO (Fei et al., (J¨ulicher et al., 1997; Kolomeisky and Fisher, 2007). The 2014)), which is the functional state in the presence of mis- monograph by Howard, written nearly twenty years ago, folded proteins. The blue and aqua colors represent the seven covers all aspects of cytoskeletal motors and is a land- fold symmetric GroEL rings. The red and the pink correspond mark in this field. It provides conceptual and practi- to the co-chaperonin GroES. (e) Structure of the DnaB he- licase with six subunits assembled as a ring. The figure was cal guidelines needed to understand the fundamentals created using the coordinates in the PDB code 4ESV (It- of motor mechanics. Questions of generality, such as sathitphaisarn et al., 2012). Structures (a)-(c) are adapted describing movement of a generic motor either in iso- from (Vale, 2003a). lation or as a collection, were addressed using princi- pally the model by Julicher, Adjari, and Prost (J¨ulicher et al., 1997), whereas Kolomeisky and myosin V share some structural similarities. and Fisher (Kolomeisky and Fisher, 2007) summarized the development and practical applications of stochas- tic kinetic models. Here, our focus is on the more re- A. Molecular motors cent developments and applications of physical princi- ples that have started to provide a unified perspective Let us first consider the three motors whose structures for not only molecular motors but a large of class of are schematically shown in Fig. 3 (a-c). Although there machines with vastly different functions. These new in- are differences between them, the parts list is roughly sights have become possible by expanding the scope of the same. These motors are dimeric. The nucleotide traditional stochastic chemical kinetics models for mo- binding sites are in the two motor heads, which are con- tors, helicases, and molecular chaperones, incorporation nected through a mobile linker (the lever arm in myosin) of polymer physics concepts to account for the architec- to a tail domain involved in dimerization and cargo- ture of motors, and use of coarse-grained models in sim- binding. Changes due to nucleotide binding and hydrol- ulating the stepping kinetics of motors and the dynamics ysis in the motor heads result in conformational changes of large scale allosteric transitions. in the linker that propel the motor on the cytoskele- tal filaments. For the motors to be processive, which means they take multiple steps before disengaging from II. STRUCTURES the track, one head has to be bound till the detached head rebinds to a site on the track. This involves communi- The structures of the biological machines that we con- cation between the heads and is referred to as gating, sider in our perspective (kinesin, myosin, dynein, the the origin of which is still not fully understood at the E. Coli. chaperonin machinery GroEL/GroES, and he- molecular level. licases) are shown in Fig. 3. Inspection of the figures Kinesin: There are at least forty five members be- shows that there are considerable variations in the ar- longing to the kinesin superfamily in mouse and human chitectures, although kinesin-1 (or conventional kinesin) genomes (Hirokawa et al., 2009). They are all micro- 5 tubule (MT) bound motors, which walk unidirectionally Chaperonins: The beautiful and unusual structure towards the plus end of the MT (for example kinesin-1) with seven fold symmetry of the E. Coli. chaperonin or to the minus end (Ncd motor). Kinesin-1 takes pre- machinery, consisting of a complex between GroEL and cisely 8.2 nm steps, which corresponds to the distance GroES, resembles an American football (Fig. 3d). The between two adjacent α/β dimers, which are the GroEL/GroES machine helps in the folding of recalci- building blocks of the MT. The size of kinesin-1 is a few trant proteins that do not fold spontaneously. The struc- nanometers whereas the length of the stalk is in the range ture in Fig. 3d (reported in (Fei et al., 2014)) is the func- of 30-40 nm. The kinesin-1 velocity depends on the con- tional state of this stochastic machine that is one of the centration of ATP, saturating at high values. The motor populated states during the catalytic cycle of GroEL in moves towards the plus end of the MT with maximal ve- the presence of misfolded substrate proteins (SPs). There locity of approximately 800 nm/s (Visscher et al., 1999), are two chambers in which the SPs could be sequestered. and is capable of resisting forces on the order of 7 pN The volume of each of the chambers in the structure in 3 (Carter and Cross, 2005; Visscher et al., 1999) (kBT = Fig. 3e is about 185,000 A˚ , which is about twice the vol- 4.1 pN nm where kB is the Boltzmann constant and T is ume in the relaxed structure in the absence of nucleotides the temperature).· and GroES, the co-chaperonin. The spectacular change in the chamber volume that occurs during the catalytic Myosin: The number of genes encoding for myosin mo- cycle, with accompanying alterations in the chemical na- tors in roughly thirty five (Sellers, 2000). The superfam- ture of the cavity interior, changing from hydrophobic ily of the actin bound are divided into fifteen to polar, is the mechanism of annealing action of this classes (Hartman and Spudich, 2012; Sellers, 2000). With machine (Todd et al., 1996). the exception of myosin VI all other members of this fam- Helicases: Helicases, which are involved in all aspects ily walk towards the plus end of actin. The structure of of nucleic acid metabolism (Lohman, 1992; Lohman and myosin V (Fig. 3) shows that the motor heads are con- Bjornson, 1996), function by coupling nucleoside triphos- nected to the lever arms, which are made up of six IQ phate (NTP) hydrolysis, to either translocate on single motifs. The lever arm is 36 nm-long stiff unit (the strand nucleic acids (ssNAs) or unwind double-stranded persistence length exceeds≈ 100 nm), whose size is com- (ds) DNA. Depending on their sequences they are clas- mensurate with half the helical repeat length of F-actin. sified into six super families (SFs). The structural di- The maximal velocity of myosin V, whose motor head is versity of helicases can be appreciated by noticing that larger than kinesin-1, is roughly 500 nm/s (Baker et al., sequences classified under the SF1 and SF2 families are 2004) with a stall force between 2-3 pN (Kad et al., 2008; non-ring forming whereas those in SF3-SF6 form ring Mehta et al., 1999; Uemura et al., 2004a; Veigel et al., structures. The hexametric structure (Itsathitphaisarn 2002). et al., 2012) of the replicative helices (DnaB) from bacte- Dynein: Cytoplasmic dynein, discovered over fifty ria, which does belong to the AAA+ family, is shown in years ago (Gibbons and Rowe, 1965) and pictured in Fig. Fig. 3d. However, unlike GroEL (see Fig. 3d) DnaB has 3b, walks somewhat erratically with a broad step-size the expected crystallographically allowed six fold symme- distribution (DeWitt et al., 2012; Reck-Peterson et al., try. How the NTP chemistry is coupled to translocation 2006a) on the MT towards the minus end. The struc- and unwinding remains an outstanding unsolved theoret- tural features of dynein are different when compared with ical problem. other cytoskeletal motors (see Fig. 3). First, the motor head belongs to the class of AAA+ family, which means dynein must have evolved from a different lineage com- B. Catalytic Cycle pared with myosin and kinesin. Second, other AAA+ enzymes, such as bacterial chaperonin GroEL (Fig.3d) All machines undergo a catalytic cycle, not unlike and protein degradation machines, are oligomeric assem- man-made motors, in which fuel, usually in the form of blies. In contrast, the hexameric ring that constitutes ATP, bound to a nucleotide binding site (or sites) is hy- the motor domain assembles from a single polypeptide drolyzed. These events trigger conformational changes chain! Third, the size of the dynein motor head is sig- that produce motion, which in motors and helicases re- nificantly larger than those of kinesin and myosin. The sults in stepping on the polar tracks or translocation length of the motor head of dynein along its longest axis on single stranded nucleic acids. In chaperones the nu- is about 25 nm, which is in contrast to the diameter of cleotide chemistry is linked to conformational changes, kinesin, which is only about 5 nm. Finally, although which in turn perform work on the protein or ribozyme there are six nucleotide binding sites in dynein, hydroly- to be folded. The link between ATPase cycle and func- sis in only two (perhaps three) are relevant for its motil- tion is somewhat different in GroEL, which we describe ity. The velocity of dynein at 1 mM ATP is roughly 100 below. The catalytic cycle for myosin V in the simplest nm/s (Reck-Peterson et al., 2006b) with a stall force∼ that form, which suffices for our purposes, is reproduced in is 7pN (Gennerich et al., 2007; Toba et al., 2006). Fig. 4. There are four crucial steps (see Fig. 4). In ≈ 6

2016; Gladrow et al., 2016). These are the key features characterizing the out-of-equilibrium nature of biologi- cal machines driven by non-conservative forces. Conse- quently, molecular machines can be thought of as active systems, and hence a note on diffusive motion is war- ranted. Chemical free energy released upon hydrolysis of ATP or GTP is the driving force for directed motion of molecular motors. However, if the energy source is not explicitly modeled in the description of the associated dynamics, molecular motors could be regarded as self- propelled active particles. To our knowledge such an ap- proach has not been pursued to calculate experimentally measurable quantities, such as the force-velocity relation or the distribution of run lengths of motors. The dynamics of active particles is fundamentally dif- ferent from passive particles at equilibrium or moving FIG. 4 A simple representation of the catalytic cycle of under the influence of a conservative external field. For myosin V describing the stepping of the trailing head towards example, the mean drift velocity of a spherical colloidal the plus end of actin, which is the dominant pathway in the absence of resisting force. The main text describes the de- particle with a drag coefficient γ, and charge q subject to tails. As described later in this perspective there are four a constant electric field E is VD = qE/γ. This relation other pathways that have to be accounted for in order to is readily obtained by balancing the Stokes force (γVD) produce a quantitative theory of the stepping kinetics of this and the force exerted by the field (qE). The diffusion motor. The figure is reproduced from (Vale, 2003b) constant is D = kBT/γ, which is the Stokes-Einstein re- lation. Hence, the distribution of the particle position is given by the probability density, the initial state, ATP binds to the trailing head (TH),  (x x V t)2  with ADP in the (LH). The premature release of ADP P (x, t) exp − 0 − D . (1) from the LH is slowed by rearward tension, which is an ∼ − 4Dt example of gating. Upon ATP binding to the TH the Starting from an initial position x0, the particle moves on interaction with F-actin is weakened, resulting in its de- average at velocity VD and the distribution of positions tachment from F-actin. During the diffusive search by 2 spreads with time as (δx) = 2(kBT/γ)t. Of particular the TH for the forward binding site, ATP is hydrolyzed, note is that D is definedh independentlyi of the particle producing ADP and the inorganic phosphate, P . In this i velocity VD, so that the magnitude of D is not altered by state, the TH binds to F-actin after which Pi is released increasing the field strength. Conversely, the velocity of from the new LH followed by ADP release from the TH, the particle does not depend on the ambient temperature and the cycle continues till the processive run ends. Of T . Thus, VD and D are mutually independent of one course, this simple description is incomplete because the another. rates for nucleotide binding vary depending on the nu- In stark contrast, for an active particle, whose motion cleotide concentration. Nevertheless, this simple reaction is powered by the internal fuel, the diffusivity is no longer cycle, whose main features hold for all myosin motors, is independent from the driving velocity. The effective sufficient to nearly quantitatively characterize many ex- 2 diffusion constant, Deff, defined as limt→∞ (δx) /2t, perimental observables (see below). The catalytic cycle depends on a set of non-thermal parametersh andi vio- for kinesin is similar except that the interaction between lates the FDT (Liu et al., 2011; Tailleur and Cates, the MT and kinesin is weakest if the motor head con- 2008). For transport motors exhibiting one-dimensional tains ADP. All other nucleotide states (the no nucleotide movement along the cytoskeletal filament, both V = apo, ATP bound state, the state with ADP and P ) bind 2 i limt→∞ d x(t) /dt and D = (1/2) limt→∞ d δx(t) /dt strongly to the MT. can be measuredh i directly using a given timeh tracei of a motor, which can be measured using single molecule opti- cal tweezer experiments. Indeed, such measurements on III. BIOLOGICAL MACHINES ARE ACTIVE SYSTEMS. kinesin-1 (Visscher et al., 1999) were used to obtain the velocity (V ) and diffusion coefficient (D) from the global Biological machines are non-equilibrium systems that analysis of the stepping trajectories. These values were are driven by non-conservative forces require constant used (Visscher et al., 1999) to estimate the randomness supply of energy. Therefore, it is not surprising that parameter r = 2D/d0V where d0( 8.2 nm) is the step detailed balance (DB) relation and the fluctuation- size of kinesin-1. When the diffusion≈ constant D is cal- dissipation theorem (FDT) are violated (Battle et al., culated from the randomness parameter with the knowl- 7 edge of V and d0, and D is compared with V measured ments. However, typically the number of observables are under the same conditions, it can be shown that D in- very few, making matching the predictions of the SKM creases monotonically with V (Hwang and Hyeon, 2017) to experiments is difficult. Furthermore, it is often dif- in contrast to passive diffusion. This result is a con- ficult to interpret the physical meaning of the extracted sequence of the enzyme catalytic turnover (Hwang and parameters in terms of the underlying motor architecture Hyeon, 2017). As already discussed the effective diffu- and the underlying biochemical cycle. As a consequence, sivity D for kinesin-1, which is an active particle whose it is necessary to strike a balance between the “minimal- dynamics is fueled by ATP hydrolysis free energy, does ity” of a model, which reduces the risk of over-fitting not obey the FDT. This illustration shows that these en- and increases the generality of SKMs and the predic- ergy consuming enzymes operate out of equilibrium. tive power, and the “comprehensiveness” of the network considered. This comes at some risk; for instance, one may be tempted to neglect certain transitions that are IV. STOCHASTIC KINETIC MODELS “fast” compared to others, and therefore are not expected to contribute to the overall phenomenology of the ma- In this section we discuss the development of stochas- chine. In addition, it may seem reasonable to ignore off- tic kinetic models (SKMs) as a means of describing the pathway, slow and rare transitions. On the other hand, in function of different types of molecular machines. As order to understand the function of molecular machines, stated above and shown explicitly in Fig. 4 for myosin experimentalists often probe their response to changes in V, all molecular machines go through cycles during which the environment (for instance modifying the ATP con- they hydrolyze ATP (or GTP) and perform a some func- centration), or by applying mechanical perturbations and tion. Bulk-kinetics, single-molecule experiments, and studying the response of the motor. As the environment structural studies have shown that the intermediates ex- or perturbations change, the nature of the “rate-limiting” plored by the molecular machines during the cycle have step as well as the likelihood of alternative stepping path- distinguishable structural or kinetic features and have ways might change. Therefore, a simplified description provided the rate for going from one state to the next. that ignores certain intermediate states might only work The temporal resolution in these studies is on the order for a limited set of experimental conditions, thereby re- of milliseconds [although new techniques enable the ex- ducing the number of measurements that can be used to ploration of time-scales as fast as 55 µs (Isojima et al., train the parameters or to falsify the predictions. Finally, 2016)]. It follows that the dynamics≈ of the molecular we note en passant that the determination of an appro- machine can be described as a discrete-state (the dis- priate catalytic cycle is predicated on a few experimental tinguishable intermediates should be experimentally ob- observables only, and it may not be complete enough for served), continuous-time Markov model – a framework all the states that a machine might sample during its that falls under the rubric of the “master equation.” In function. this model, the cycle of a is reduced to a network of states connected by edges that identify the available transitions between the intermediates. Thus, by solving the dynamics associated with an appropriate network it is possible, in principle, to account for exper- imental observables in terms of the underlying network Another issue with SKMs is that structural informa- parameters. tion is incorporated in the model only in a very approxi- SKMs have a number of appealing features: (i) they mate way, normally by introducing a parameter that al- have a direct connection with the biochemistry associ- lows the rate to depend on the load applied according ated with the molecular machines; (ii) the rates that to the Bell model (Bell, 1978). This reduces the capa- constitute the model are measurable; (iii) it is some- bility of the model to incorporate the wealth of experi- times possible to find an analytical solution, and (iv) mental structural information, and decreases the predic- more complex networks may be solved numerically; (v) tive power of the proposed paradigm. However, in some finally, SKMs can be directly related with thermodynam- cases a tractable analytical theory that incorporates lever ics, which makes the model instructive and appeals to the arm structural information into a kinetic model is possi- interests, and intuition of a vast scientific community. ble (Hinczewski et al., 2013), as discussed in more detail Nevertheless, SKMs also have some shortcomings: first below. Despite these limitations, the framework underly- and foremost, they often have a large number of param- ing SKMs not only provides a convenient way to analyze eters. For instance, a simple model for a molecular ma- experiments, but also has been used to address concep- chine performing mechanical work against a fixed load tual questions related to the efficiency of biological ma- and described by a single-cycle network with N interme- chines. For these reasons, the theory underlying SKMs diate states has 4N 2 independent parameters, which is an integral part of describing the function of molecular have to be determined− by fitting to appropriate experi- motors. 8

A. The Master Equation

We assume that the system is described by N distin- guishable intermediate states. The probability of being in state i at time t is given by pi(t), and the time evolution of this probability is governed by the master equation,

dpi(t) X X = p (t)w p (t)w , (2) dt j ji − i ij j j where wij 0 is the rate for the transition i j, ≥ P → subject to the constraint i pi(t) = 1. Because of mi- croscopic reversibility, if w = 0 then w = 0. As ij 6 ji 6 t , the probabilities pi(t) become independent of time.→ ∞ This stationary solution of the master equation eq describes an equilibrium system (pi ) if detailed balance holds, that is if across all the edges of the network the eq eq net flux is zero: pi wij pj wji = ∆Jij = 0. However, dp /dt = 0 is also satisfied− by the less stringent relation FIG. 5 Example of kinetic network. The network has 6 states i and 7 edges [see panel (a)]. Panel (b): the partial graphs ori- P (pssw pssw ) = P ∆J = 0. Under these con- j i ij − j ji j ij ented in order to extract the stationary probability of state ditions the system could be in a non-equilibrium steady 1. Panel (c): There are three cycle fluxes, F (upper three state (NESS). An isolated system is expected to reach figures), B (bottom three), and D (central figure). From the an equilibrium state, whereas coupling with an external figure we extract the values of Σν in Eq. 4 for the three cy- energy source enables the creation and persistence of a cles. These are given by the product of the rates flowing to- NESS. wards the cycle [see T. L. Hill (Hill, 2005a) for details]: ΣF = From a mathematical standpoint, given a kinetic net- w32w45 + w43w32 + w34w45;ΣB = w12w65 + w61w12 + w16w65; Σ = 1. work described by the master equation, the stationary D ss eq probabilities pi (or pi ) can be obtained using graph- theoretical arguments. The details may be found in the where Aν (or βAν ) has been termed the action func- works of Hill (Hill, 1966, 2005a; Hill and Chen, 1975). tional (Lebowitz and Spohn, 1999) or affinity (Schnaken- Briefly, one may construct the set of partial diagrams berg, 1976) of the cycle ν. Note that the net direc- such that all the states are visited but no cycles are tion of completion of cycle ν is given by the sign of formed (see Fig. 5b). The stationary probability of be- ∆Jν = Jν+ Jν− . It is easy to show that (Hill and ing in state i is proportional to the sum of all the par- Simmons, 1976),− tial graphs oriented in such a way that the fluxes con- verge towards state i, or σi. The normalization factor is ∆Jν Aν 0. (6) P ≥ Σ = i σi, so that the stationary probability is, The above equality holds only when Jν,+ = Jν,−. In ss σi pi = . (3) other words, ∆Jν and Aν have the same sign, which Σ means that the value of the affinity dictates the direc- The stationary flux along one edge may be written as tion of the cycle (Hill and Simmons, 1976). a sum of contributions from all the cycle fluxes of the In order to identify the physical interpretation of system. Each cycle can be completed in two directions: these mathematical identities we need to connect the one, counterclockwise, labeled as “+”; the other, clock- rates with thermodynamic quantities, such as energy, wise, termed “-”. The cycle fluxes in the “+” and “-” entropy, and the chemical potential. Following three directions are given by the following relationships (Hill, different approaches we show that the affinity (Eq. 5) 2005a), is related to the entropy produced during a cycle. The first two strategies involve the Shannon entropy (Liepelt Σν Πν± J ± = , (4) ν Σ and Lipowsky, 2007b; Lipowsky and Liepelt, 2008; Schnakenberg, 1976) and fluctuation theorems (Crooks, where Σν is a combination of rates that are specific for 1998; Seifert, 2005a,b, 2012), and they will be dis- cycle ν,Πν± is the products of the rates of the cycle per- cussed without specifically referring to the experimental formed in the “+” (Πν+ ) or “-” (Πν− ) direction, and the hallmarks for motor velocity described in the previous denominator Σ is defined above (see Fig. 5c). It follows sections. The last method, which is based on estab- that, lishing a connection between pseudo-first-order rate J + Π + constants and the chemical potential (Hill, 2005a), will ν = ν = eβAν , (5) Jν− Πν− be developed in the context of molecular motors. 9

At equilibrium, the rate of hydrolyzing and synthesizing ATP is the same. However, this is not what happens in a cell, where B. Thermodynamics the concentration of ATP is maintained under homeo- static control (Wang et al., 2017) far away from equilib- Molecular machines are enzymes (E) that catalyze the rium; typical concentrations are [ATP] 1mM, [ADP] chemical transformation of substrate molecules (S P ), 10µM, [P ] 1mM (Howard, 2001), resulting≈ in a chem-≈ → i ≈ which in general can be described by the Michaelis- ical potential X = ∆µhyd of 25 (Howard, 2001), Menten kinetics (E + S ES E + P ). Of course, which makes ATP− hydrolysis a≈ spontaneous reaction → the process is reversible, and the enzyme also catalyzes (∆µhyd < 0). This means that the enzymatic cycle of the the reverse reaction – the transformation of P into S. molecular machine will be driven in the direction that In the case of molecular machines, the substrate is ATP, consumes ATP, and ATP hydrolysis provides the driv- and the products of ATP hydrolysis are ADP and or- ing force that enables the molecular machine to perform thophosphate (Pi). (Some molecular machines catalyze work. Without accounting explicitly for the whole cellu- the hydrolysis of GTP; mutatis mutandis our considera- lar apparatus involved in dictating and maintaining the tions do not change.) set-point ATP level, in the simplest theoretical model the We imagine the following experimental setup: a molec- system is assumed to be in contact with some devices re- ular machine operates in a solution containing ATP, ferred to as “chemostats” capable of maintaining the ini- ADP, and Pi. The chemical potentials of these three tial concentrations of ATP, ADP, and Pi (Lipowsky and species are given by (Hill, 2005b), Liepelt, 2008; Lipowsky et al., 2009; Qian and Beard, 2005; Seifert, 2011b). After each cycle the chemostat 0 µATP = µATP + kBT ln [ATP] removes the products from solution and replenishes the 0 substrates, thereby ensuring that after each enzymatic µADP = µADP + kBT ln [ADP] cycle the initial condition is reset, which enables the cre- µ = µ0 + k T ln [P ] Pi Pi B i (7) ation of a NESS for t . → ∞ X = ∆µhyd = µATP µADP µPi = The system is also in contact with a thermal reser- − − − Keq[ATP] voir with which it exchanges heat (Lipowsky and Liepelt, = kBT ln , 2008; Lipowsky et al., 2009; Seifert, 2011b). We assume [ADP][Pi] that the thermal reservoir operates “quasi-statically” on 0 the time-scale of the heat exchange, and the combination where µc (µc ) is the (standard) chemical potential of species c,[c] is the concentration in solution, and Keq of the system and the thermal reservoir is energetically 4.9 105 M (Howard, 2001) is the equilibrium constant≈ isolated. Because after a cycle the motor and the solu- for ATP· hydrolysis. (Note that we should use activities tion have not changed, the entropy produced is equal to ac in Eq. 7 instead of concentrations [c]; throughout the the ratio between the heat absorbed by the thermal reser- review we will assume that a [c].) At equilibrium, voir, Q, and the constant temperature T . We exclude pV c ≈ ([ADP][Pi]/[ATP])Eq = Keq, and ∆µhyd = 0. On the work from our formulation, but we include the possibility other hand, if [ATP]K [ADP][P ], then ∆µ < 0, that the molecular machine performs work against a fixed eq  i hyd and the hydrolysis of ATP is a spontaneous reaction. load, so that for a forward (backward) step Wmech = fd0 We exclude un-catalyzed hydrolysis/synthesis of ATP, as (Wmech = fd0), where the choice of the sign implies − it occurs over time-scales beyond our interest (Hulett, that a positive force opposes forward movement. Note 1970); as a consequence, in the absence of molecular ma- that the load f is assumed to be clamped, so at every chines the concentrations [ATP], [ADP], and [Pi] are con- step, regardless of the position of the motor, the cycle is stant. The presence of the molecular machine does not repeated under identical conditions. alter the equilibrium features of ATP hydrolysis (namely, The energy X = ∆µhyd > 0 consumed during a cycle − K ), however it accelerates the rate of ATP synthe- is partitioned into work performed ( Wmech) and heat eq − sis/hydrolysis. Let the initial concentrations of ATP, released ( Q), that is, − ADP, and P make ATP hydrolysis a spontaneous re- i ∆µ + Q + W = 0, (8) action. After each catalytic cycle the enzyme attains hyd mech the same conformation that it had at the start. How- implying that the total change of entropy of the system ever, the solution conditions have changed as a substrate plus environment over one cycle is given by, (ATP) has been consumed and products (ADP and P ) i Q ∆µ fd have been created. Thus, in the presence of the molecular ∆S = = − hyd − 0 0. (9) machine the solution approaches the equilibrium ratio of T T ≥ concentrations of ATP, ADP, and Pi, and monitoring the (For backward steps, the sign in front of fd0 is the oppo- dynamics under these conditions corresponds to study- site). The inequality is due to the second law of thermo- ing the time-dependent relaxation towards equilibrium. dynamics, which states that we should to find an increase 10 in entropy during the cycle because the combination of Consider now a cyclic network of N states; the only system and environment is isolated. Eq. (9) leads to the transitions allowed from state i are i i + 1 and i observation that the maximum amount of force that the i 1. (The cyclic nature of the network implies periodic motor can resist is given by, boundary− conditions, and so state i = 0 is the same as P state i = N). In a NESS, the condition Jij = 0 ∆µhyd j fmax = − . (10) becomes ∆J = ∆J , so that the net flux across d i,i+1 i−1,i 0 all the edges is the same and equal to the net cycle flux, A number of authors have discussed variations of ∆J = J+ J−. It follows that, this thermodynamic framework. In particular, Seifert − N−1 elucidated the contribution of the thermodynamic diS Y wi,i+1 ∆JA = k ∆J ln = (15) contribution of chemostats (Seifert, 2011b); Qian and dt B w T i=0 i+1,i Beard discussed a network of reactions in which some metabolites are clamped, while other are introduced at where wi,i+1 and wi+1,i are the forward and backward rates, respectively. The term A = k T ln QN−1 wi,i+1 is a constant rate (Qian and Beard, 2005). B i=0 wi+1,i the affinity (see Eq. 5), which Hill refers to as a ther- modynamic force driving the system out of equilibrium and imposing a NESS (Hill, 2005a). Liepelt and Lipowky C. Rate of Entropy Production identified the thermodynamic force with the entropy pro- duction (Liepelt and Lipowsky, 2007b); if the rate of Let the entropy be, completing a cycle is ∆J > 0 (which implies that the cy- X S = k p ln p , (11) cle is preferentially completed in the forward, or “+”, di- − B i i i rection), one can identify the entropy ∆S produced over one cycle as, where the sum extends over all the available conforma- N−1 tions of the system. Let us take a time derivative of T diS Y wi,i+1 T ∆S = = k T ln = A 0. (16) S, and after imposing the conservation of probability ∆J dt B w ≥ P i=0 i+1,i ( i dpi/dt = 0) condition, and plugging in the master equation we get, The argument can be extended to the case in which the dS kinetic network is characterized by multiple cycles (Lie- X X −1 = kB (piwij pjwji) ln pi. (12) pelt and Lipowsky, 2007b); if ∆J is the average time dt − ν i j for completing a cycle ν, and ∆Sν is the entropy pro- Note that in a NESS, P (p w p w ) = P J = duced with that cycle, we can write, j i ij − j ji j ij 0, and as a consequence dS/dt = 0, as expected. We diS X = J ∆S 0. (17) first symmetrize the result with respect to the indexes i dt ν ν ≥ and j, and then add and subtract 1/2k P P (p w ν B i j i ij − pjwji) ln(wij/wji), which leads to (Lipowsky and Liepelt, Equation 6 indicates that each term in the summation is 2008; Schnakenberg, 1976), non-negative (Hill and Simmons, 1976). To summarize, in this section we showed how the dS 1 X X piwij 1 X X wij = k J ln k J ln . action functional (or affinity) can be identified with the dt B 2 ij p w − B 2 ij w i j j ji i j ji entropy produced. We now present a different argument (13) based upon fluctuation theorems leading to the same The first of the two terms on the r.h.s. of Eq. 13 is always conclusions. non-negative, and it has been identified as the rate of en- tropy production, diS/dt (Lipowsky and Liepelt, 2008; Schnakenberg, 1976). We note parenthetically that some authors (Hill and Simmons, 1976) do not include the fac- D. Fluctuation Theorems tor 1/2 as they interpret the summation to be carried out over the edges of the network, and not over the states. Let the probability density of taking a forward path of n steps be (i , t ; i , t ; ; i , τ), in which the system Because in a NESS dS/dt = 0, it follows that the second PF 0 0 1 1 ··· n term of the r.h.s of Eq. 13 is the rate of entropy out-flux, is in state i0 at t0, transitions to state i1 at t1 and so on and, until it reaches state in at time τ (see Fig. 6a). We use the subscript “F ” to indicate that the transitions are diS 1 X X piwij = k J ln = completed forward in time. Using the Markov property, dt B 2 ij p w i j j ji we can rewrite this joint probability as (Crooks, 1998), (14) 1 X X wij = k J ln 0. F (i0, t0; i1, t1; ; in, τ) = B 2 ij w ≥ P ··· (18) i j ji = p (t )P (i , t i , t ) P (i , τ i − , t − ). i0 0 1 1| 0 0 ··· n | n 1 n 1 11

ss pin and because each path could be performed forward or backward in time, the normalization factor = ˜ . Following Seifert (Seifert, 2005b), we define R toN be theN logarithm of the ratio between the probability for for- ward and backward paths, and for the sake of simplicity we consider only cyclical pathways, in which i0 = in. Therefore,

n−1 F Y wiα,iα+1 R = ln P = ln . (21) w PB α=0 iα+1,iα Note that the time-dependent terms in the joint proba- bilities cancel out, so we neglect them in the following. (In alternative, we may consider the time at which these jumps occur to be fixed.) Here, R has a structure anal- ogous to that of the affinities introduced in the previous sections. More precisely, given a set of states visited we FIG. 6 Path in state space executed forward (a) and back- can rewrite, ward (b) in time. The starting point is the dot, waiting times X at a fixed state are in black, transitions are in blue. R = nν βAν , (22) ν

where nν is the net number of times the cycle ν has Assuming that the system is in a steady state at t = been completed during the n-step path associated with ss t0, we replace pi0 = pi0 . The conditional probabili- R. In order to justify this expression, note that every −Wi(tj −ti) ties are given by P (j, tj i, ti) = wije , where time that during the path a cycle is not completed, the W = P w . The conditional| probability P (j, t i, t ) path re-traces it self and those branch-like excursions do i j ij j| i is the product of two probabilities. The first is wij/Wi, not contribute to R. If we average over all paths that the probability of making a transition to state j among complete a cycle of length n, we find the following se- all the other possibilities when starting in state i. The quence of identities (Seifert, 2005b), −Wi(tj −ti) second is Wie , the probability that this transi- X X e−R = P e−R = P = 1. (23) tion happens after time (tj ti). The reason for the Wi h i F B in the exponent is that the− mean waiting time to transi- paths paths −1 tion from state i to any other state is Wi , independent Using Jensen’s inequality, e−R e−hRi, leads to, of state j. This independence is a well-known property h i ≥ of escape times in Markov networks, which may be com- R 0. (24) h i ≥ puted through well-known methods (Van Kampen, 2007). From Eq. 22 we obtain, Therefore, the joint probability density becomes (Seifert, X 2012), R = n βA 0. (25) h i h ν i ν ≥ ν F (i0, t0; i1, t1; ; in, τ) = P ··· We now introduce an Arrhenius-type relationship be- n−1 (19) tween the rates, w /w = e−β∆Fij , where ∆F = −1 ss Y −Wiα (tα+1−tα) ij ji ij = pi0 wiα,iα+1 e , N Fj Fi is the free energy difference between state i and α=0 state− j. The free energy difference accounts for three where tn = τ and (t0, τ) is a normalization factor to contributions: (i) the intrinsic free energy of a state may P N R τ R τ R τ ensure that dt1 dt2 dtn F = 1. change; (ii) the motor may bind/release ATP, ADP, and i0,i1,·,in t0 t1 ··· tn−1 P The probability of completing the time-reversed path Pi; (iii) the motor may perform work against an exter- (Fig. 6b) is, nal load. After a cycle, the motor returns to the initial state, and as a consequence the intrinsic free energy does (i , τ,˜ i − , t˜ − ; ; i , t˜ ) = not carry any contribution to the cycle. The hydrolysis PB n n 1 n 1 ··· 0 0 N−1 of ATP contributed with the release of energy equal to − Y − ˜ −˜ (20) = ˜ 1pss w e Wiα (tα tα+1), ∆µhyd, and the work performed per each displacement N in iα+1,iα − α=0 of size d0 against a load f is equal to fd0. It follows that (Lipowsky and Liepelt, 2008), − in which we started in state in at time t˜n = τ tn and ∆Sν − Y kij −nν,ATPβ∆µh−βflν retrace the forward path until we reach state i0 at time = e = e kB , (26) kji t˜ = τ t . Note that the initial probability is now |i,ji∈ν 0 − 0 12 where, following the notation of Lipowsky and Lie- forward while the others hold tight onto the CF. From pelt (Lipowsky and Liepelt, 2008), i, j corresponds to a here on, we focus our discussion exclusively on dimeric directed edge from state i to j, and| thei sum is extended processive motors made of two identical enzymes of the over all the directed edges belonging to cycle ν. Here, same family, which we refer to as “heads.” With this nν,ATP is the net number of ATP hydrolyzed during cy- we may identify the leading head (LH) as the one that cle ν, and lν is the net displacement along the track (a is in front of the dimeric complex, while the other is multiple of d0). The last identity follows from Eq. 9, and the trailing head (TH). We reserve the word “motor” from Eq. 25 gives, for the motile construct, which is a dimer in the case of myosin V (Mehta et al., 1999), VI (Rock et al., 2001), X k R = n ∆S 0. (27) and X (Sun et al., 2010), conventional kinesin (kinesin- Bh i h ν i ν ≥ ν 1) (Howard et al., 1989), and cytoplasmic dynein (Reck- Peterson et al., 2006b). In the well-accepted hand-over-hand stepping mecha- V. MOLECULAR MOTORS – MODELS WITHOUT ¨ DETACHMENT nism (Okten et al., 2004; Sun et al., 2010; Toba et al., 2006; Yildiz et al., 2003, 2004a,b), the TH of the motor In this section we develop models of molecular motors detaches from the filament, overcomes the bound head, with increasing complexity, from one-state models, and advances the motor by reaching a forward target to uni-cyclic models, ending with multi-cycle kinetic binding site (TBS). After the step is completed, the role networks. We show that experimental evidence and of the two heads is reversed, and the process is repeated thermodynamic insights suggest that the introduction identically, until the motor detaches from the CF. A few of multiple cycles creates models that are physically key ingredients are necessary for processive, hand-over- more sensible. We discuss more in detail the efficiency hand movement to occur. First, in order to prevent pre- of the motors, and we conclude the section by describing mature end of the processive run, the head that remains models that account for the detachment of the molecular bound to the track during a step should not dissociate motor from the track. from the CF at least until the detached head reaches the TBS. Second, the two heads should go through their cat- alytic cycles out-of-phase (Block, 2007), so that the TH is more likely to detach from CF than the LH. This re- A. Molecular Motors are Processive Enzymes quirement necessarily implies that the two heads ought to communicate with one another through action at a The catalytic cycle of kinesin, dynein, and myosin pro- distance, which may be viewed as a complex form of al- ceeds via multiple steps, in which ATP binding, hydroly- lostery (Thirumalai et al., 2019). Third, during a step sis, and release of ADP and Pi lead to substantial changes a combination of conformational transitions in the two in the motor conformation. The presence of the cy- heads biases the diffusive motion of the free head towards toskeletal filament, CF (F-actin in the case of myosin, the TBS unless a large resistive force is applied. microtubules for dynein and kinesin), increases the rate Meeting the first two conditions requires specific fea- of ATP hydrolysis (De la Cruz et al., 2001, 1999; Hack- tures of the hydrolytic cycles that the two heads must ney, 1988; Homma and Ikebe, 2005; Ori-McKenney et al., undergo. For instance, if the rate-limiting step occurs 2010). When the structural changes are properly recti- in a CF-bound conformation, the chances of prematurely fied under the specially designed interactions with the ending the processive run are diminished. In addition, appropriate CFs, the conformational fluctuations in M the interaction between the two heads is believed to be are transduced to a predominantly uni-directional motion key in ensuring that the TH steps first, and that the along the track, and generates mechanical forces against LH is unlikely to detach during a step. This is possi- an external load. The free energy source necessary to ble if the interplay between the two heads slows down perform this movement (or work) is provided by the hy- (or “gates”) some of the steps of the LH catalytic cycle, drolysis of ATP. or accelerates them in the trailing head (also referred to Enzymes that work in conjunction with substrates (as as “gating”) (Block, 2007; Hancock, 2016; Sweeney and is the case for Ms and CFs) are said to be “processive” Houdusse, 2010). The structural bias towards the TBS is if they perform multiple catalytic cycles without fully provided by a conformational transition of the CF-bound disengaging from the CFs (Schnitzer and Block, 1995). motor that projects forward the stepping head. For Processive motors move long distances along the CFs by the three classes of motors mentioned before, the lever hydrolyzing one ATP molecule per step without dissoci- arm swing in myosin, the neck-linker docking in kinesin, ating from the CF. In general (although with some fasci- and the bent straight transition of the linker domain in nating exceptions (Inoue et al., 2002; Post et al., 2002)) dynein provide→ the requisite forward bias. It is remark- processive movement requires the cooperation of mul- able that these structural transitions, induced by ATP tiple enzymes: some members of the ensemble proceed binding and hydrolysis, which occur on relatively small 13 length scales are amplified by responses in other struc- tal value for the Michaelis constant is KM & 50 µM, tural elements. The terms “power stroke” (Dominguez and for [ATP] & 1 mM the velocity of kinesin is satu- et al., 1998) and “Brownian ratchet” (Rice et al., 1999) rated to its maximum value Vmax = d0kcat 8 nm/10 have been used to schematize this forward bias, depend- ms = 0.8 µm/s (Kolomeisky and Fisher, 2007;≈ Visscher ing on whether the emphasis is placed on the mechanistic et al., 1999). Similar single-molecule studies have been (order order) or stochastic (disorder order) nature of conducted on other motors, showing that the Michealis- the conformational→ transitions. → Menten scheme provides a good paradigm to rationalize In the simplest cases, the step-size of a motor is the ATP concentration dependence of velocity. However, commensurate with the filament repeat – for motors the parameters of the fit depend on the motor: for myosin moving along the MT, d0 8.2 nm, whereas d0 36 nm V it was found that Vmax 0.55 µm/s, with KM 163 µ for actin-based motors (myosins).≈ However,≈ not all M (Baker et al., 2004), whereas≈ the slower myosin≈ VI has motors walk precisely by following this rule: myosin Vmax 0.16 µm/s and KM 274 µM [fit of Eq. 28 to VI and dynein, for instance, display a broad step-size data from≈ Elting et al. (2011)].≈ distribution, in which hand-over-hand steps are inter- Experiments employing have shown twined with inchworm-like movements and frequent that external loads affect the ATP chemistry in the cat- backward steps (Nishikawa et al., 2010; Reck-Peterson alytic site as well as the motility of the motor (Block et al., 2006a). Of course, the variability in the step et al., 2003; Oguchi et al., 2008; Veigel et al., 2005; Viss- sizes is also a consequence of the architecture of the cher et al., 1999). Thus, the resulting force-velocity-ATP motor. Thus, both the structural design of the motor relationship has been a landmark measurement, which is and the coupling to catalytic cycles determine not only used as a constraint to decipher the mechanism under- the stepping kinetics but also the precision of the step lying kinesin motility, and to construct the appropriate sizes. SKMs. Incorporation of the effect of load into MM model was suggested by making the parameters kcat and KM force-dependent (Schnitzer et al., 2000),

B. Processive Motor Velocity d k (F )[ATP] V (F, [ATP]) = 0 cat (29) kcat(F )/kb(F ) + [ATP] A variety of single-molecule techniques are able to inform on the motile properties of molecular motors. o F δcat/kB T where kcat(F ) = kcat/(pcat + (1 pcat)e ) and The motor or the track may be labeled with a fluores- o F δb/k−B T kb(F ) = kb /(pb + (1 pb)e ). The fit of the cent (Nishikawa et al., 2010; Yildiz et al., 2003) or re- F -dependent velocity− data of motor using Eq.29 is fractive (Isojima et al., 2016; Mickolajczyk et al., 2015) reasonable as long as the magnitude of load (F ) is small probe. Monitoring the time-dependent changes in the (Schnitzer et al., 2000). However, the conventional MM location of the such probes enables the determination model has an intrinsic drawback when the external load of the velocity of the motor as a function of nucleotide approaches the stall force value ( 7 pN for kinesin) concentration. Alternatively, optical trapping techniques and becomes greater than the stall≈ force. An increase may be used to follow the displacement of the motor or of load ought to induce backward stepping (V < 0); yet the filament, and to investigate the motor response to no modification of Eq.28 produces a negative velocity! the external load (Block et al., 1990; Finer et al., 1994; One possible fix for the failure of the MM model at Howard et al., 1989). Much of our current knowledge large load is to permit in the model a reverse reaction about kinesin comes from single molecule experiments current, which is realized by rendering every elementary (Block et al., 1990; Howard et al., 1989), which were de- reaction step within the kinesin cycle “reversible.” We veloped approximately at the time of the discovery of ki- discuss later how such models have been constructed in nesin (Brady, 1985; Vale et al., 1985). It has been shown the context of Markov jump processes. that each step of kinesin motors along MTs is tightly coupled with hydrolysis of one ATP molecule, that is, a single event of ATP hydrolysis by kinesin leads to 8 nm step along MTs (Schnitzer and Block, 1997). Thus,∼ C. Periodic Lattice Model within the Michaelis Menten (MM) scheme, the velocity of kinesin is associated with enzyme turnover rate is, In a molecular motor (i) binding, release, and chemi-

kcat[ATP] cal transformations of ATP, ADP and Pi from the mo- V = d0 , (28) tor head domain, and (ii) the advancement along the KM + [ATP] track occur over much slower times (ms - s) compared where d0 = 8 nm is the step size of the kinesin, kcat to the fluctuations of molecular conformations (ns - µs). and KM are the rate of catalysis and the Michaelis con- The time-scale separation enables the construction of a stant, respectively. For kinesin-1, the typical experimen- Markov jump processes in which the state of the system 14 is defined by two coordinates: (i) one of the N intermedi- u = u*[ATP] ates defining the chemical state of the substrate, and (ii) the location l along the track. The probability density ⋯1l−1 1l 1l+1⋯ of the i-th state (i = 1, 2,...,N) at filament site l obeys w w⇌ ⇌ the master equation, = *[ADP][Pi] (a) l+1 u u u dPi(l, t) X X 0 N 1 = 1*[ATP] = [wj,l→i,l0 Pj(l , t) wi,l→j,l0 Pi(l, t)] 1 dt − ⇌w l0=l−1 j w = w*[ADP][P ] 2 1 ⇌1 i (30) N 2 P P 0 uN w w where l i Pi(l, t) = 1, and wj,l→i,l is the rate for go- −1 N 3 u2 ⇌ ing from state j in location l to state i on F site l0 = l, l 1. ⇌ ± Under the assumption that the track is infinite and pe- N w riodic, that is that the rates do not depend on l, one − 1 4 3 ⋯ u may sum the r.h.s and the l.h.s of the master equation 4 ⇌3 and eliminate l. As a result, we obtain the master equa- (b) tion in Eq. 2, and we can use the theoretical framework presented before. FIG. 7 Example of simple kinetic networks. Panel (a): the Periodic one-dimensional models, which were orig- network has only one statte, N = 1. The forward rate is given by u = u∗[ATP], the backward rate is w = w∗[ADP][P ]. Both inally suggested by Derrida (Derrida, 1983) in order i forward and backward rates are pseudo-first-order, with units to study the mean velocity and diffusion constant of mM−1s−1 and mM−2s−1, respectively. Panel (b): multi-state random systems, have been widely adopted in describing model. Here, the rates u1 and w1 are arbitrarily chosen to the dynamics of molecular motors. In particular, depend on [ATP] and [ADP][Pi], respectively. Panel (c): a Fisher and Kolomeisky (Fisher and Kolomeisky, 1999a, cycle corresponding to the multi-state model. 2001; Kolomeisky and Fisher, 2007) have pioneered this approach and specifically used the reversible ki- netic model to describe the motility of kinesin-1 and backward rates is given by, myosin V. Numerous other studies have used vari- ∗ u u [ATP] βX ants of these models to discuss the thermodynamic (a) = ∗ = e features of stochastic motors. We review the main w w [ADP][Pi] ∗ (32) successes and shortcomings of these models in the con- u (b) = Keq, text of molecular motors as their complexity is increased. w∗

where we recall that X = ∆µhyd. The average, station- ary velocity is given by the− step-size times the difference between the forward and backward rates,   1. One-state Models v = d (u w) = d w eβX 1 . (33) 0 − 0 − In the simplest possible model, the motor is defined by The crucial insight from this expression pertains to a single (N = 1) chemical state at each site along the the interplay between the non-equilibrium steady-state track (Fig. 7a). The rate of forward stepping, u, depends (NESS) homeostatically maintained by the cell and on the concentration of ATP. Microscopic reversibility molecular motor function: in the cytosol X > 0, thus demands that the backward transition is possible as well, v > 0; at equilibrium, X = ∆µhyd = 0, implying that which is then nominally associated with ATP synthesis. v = 0. The rates of forward and backward stepping are modeled In the absence of applied load, the motor moves with- as pseudo-first order processes, out performing any work. In this case, the entire free en- ergy extracted from the hydrolysis of ATP is dissipated ∗ ∗ u = u [ATP], w = w [ADP][Pi], (31) into heat, Q = ∆µhyd. From Eq. 15, the rate of heat dissipation per step− is given by, ∗ ∗ leading to u/w = u /w [ATP]/([ADP][Pi]). In equi- d S u · T i = Q˙ = k T (u v) ln = (u v)( ∆µ ) 0. librium, the probability of going forward is identical to dt B − v − − hyd ≥ the probability of backward stepping, which implies that (34) (u/w) = 1. From Eq. 7, at equilibrium ∆µ = 0 Clearly, Q˙ 0, and the equality holds in equilibrium, Eq hyd ≥ and ([ADP][Pi]/[ATP])Eq = Keq. As a consequence, us- where u = v. If instead the motor is subjected to a ing Eq. 7 we obtain that the ratio between forward and force, f, opposing its movement, at each step the motor 15 performs mechanical work W = fd0. In this case, the free is zero when f = 0 and at the stall force (where v = 0), ∗ energy released by ATP hydrolysis is partitioned between thus for some force 0 < f < Fs it reaches a maximum ∗ work and heat, Q = ∆µhyd W . In order to account for P . The force-velocity curve, power output, efficiency, the presence of applied− load,− the rates may be modified and efficiency at maximum power (η∗) have been used to using the Bell model (Fisher and Kolomeisky, 1999b), compare ratchet-like motors (θ = 1) with power-stroke- driven (θ = 0). Given a value of the driving force X, ∗ −βθfd (a) u(F ) = u [ATP]e 0 , power strokes result in larger velocity at fixed F , and ∗ β(1−θ)fd0 display higher efficiencies and exert more power at a given (b) w(F ) = w [ADP][Pi]e . (35) u velocity (Wagoner and Dill, 2016). Furthermore, both (c) = eβ(X−fd0) ∗ ∗ w η and P at fixed X increase as θ 0 (Schmiedl and Seifert, 2008). Finally, from Eq. 15 we→ write, Here, the coefficient θ indicates how the load is parti- tioned between the forward and backward steps, and it diS u T = Q˙ = (u w) ln = (u w)( ∆µ fd ), identifies the position of the transition state along the dt − w − − hyd − 0 direction of motion (Fisher and Kolomeisky, 2001). For (39) θ = 0, the forward step is unaffected by force, whereas where we used Eq. 35c and Eq. 36. Note that, again, Q˙ 0, and the equality holds at equilibrium (u = w). backward steps are insensitive to load if θ = 1. These two ≥ extreme cases have been referred to as “power stroke” One-state models are appealing because of their and “ratchet”, respectively (Howard, 2006). In a “power simplicity; for instance, they have only two fitting pa- stroke”, the chemical step (crossing the transition state) rameters with straightforward physical interpretations: ∗ occurs before the movement forward of the motor. In w sets the time-scale (see Eq. 33), and θ establishes the contrast, in the case of a “ratchet” the motor fluctuates location of the transition state during a step. However, between the sites l and l + 1 until it is captured in the these models fail in reproducing the dependence of the forward site by ATP hydrolysis. In the presence of back- velocity on ATP concentration observed in experiments ward load the velocity becomes, (Eq. 28). According to Eq. 33 v grows without saturat- ing as [ATP] increases, which is not physical. In order   β(1−θ)fd0 βX−βfd0 v = d0w(0)e e 1 . (36) to solve this problem, one must use kinetic models with − N > 1. It is clear that the motor stalls under a resistive load of,

X ∆µhyd fstall = = , (37) d0 − d0 2. Multi-state, Uni-cycle Models at which the mean motor velocity becomes v = 0. It is Figure 7b shows a kinetic model for molecular motors clear that fstall = fmax for this model. Using Eq. 37 to- with N > 1 intermediates. Assuming that the filament is gether with the value of ∆µhyd in the cell and the average periodic, we can drop the label l and construct an equiv- step-size of a motor, one can predict the value of fstall: for alent uni-cycle in Fig. 7c, where the rate uN is associated kinesin fstall 12.5pN, for myosin V fstall 2.8nm. In ≈ ≈ with a forward step, and w1 with a backward displace- the case of myosin V, fstall approximates the value of the ment. From Eq. 4 we obtain the following relationships, stall force measured experimentally; for kinesin, the fmax is about twice as large as the stall force. Note also that QN u QN w (a) J + = i=1 i ,J − = i=1 i , according to Eq. 37, Fs depends on [ATP]; a few experi- Σ( u, w ) Σ( u, w ) ments have suggested that this is not the case for a vari- { } { } QN u J + ety of motors (Carter and Cross, 2005; Gennerich et al., (b) i=1 i = = eβX , QN J − 2007; Nishiyama et al., 2002; Toba et al., 2006; Uemura j=1 wj (40) et al., 2004a), although other studies have observed such ∗ QN u ui adependence (Visscher et al., 1999). The independence (c) 1 i>1 = K , ∗ QN eq of Fs on [ATP] might be reasonable because each head w1 j>1 wj has only one binding pocket for ATP, and hence [ATP] should not determine Fs. The relationship v(f, ∆µ) al- lows one also to compute the power output of the motor, where J + and J − are the fluxes to complete a cycle defined as P = vF , and the motor efficiency, in the clockwise and counterclockwise direction, respec- tively, and Σ( u, w ) is a function of all the rates (see fd η(f, X) = 0 , (38) Eq. 3). The connection{ } with thermodynamics is provided X by Eq. 40b, which is equivalent to Eq. 32a; Eq. 40c holds that is the ratio of the work produced over the energy because in equilibrium X = 0 (see Eq. 32b for the N = 1 released by the hydrolysis of ATP. Note that the power case). For the case of uni-cyclic network models, the net 16 steady state flux (∆J = J + J −) along the cycle is ob- Following Eq. 15, the rate of heat released during a − ss ss tained by calculating ∆J = wi,i+1pi wi+1,ipi+1 at any cycle (Hill, 2005a; Qian, 2007; Qian and Beard, 2005; edge between the two neighboring chemical− states along Toyabe et al., 2010; Zimmermann and Seifert, 2012) is, the cycle. Therefore, the velocity, which is obtained as uN + the flux across edge N 1, is given by, J  Q˙ = ∆JkBT ln = ∆J( ∆µhydr fd0) = E˙ +W˙ 0, w1 J − − − ≥ (46) QN ˙ − βX i=1 wi βX where E is the rate of free energy expended per cy- v = d0∆J = d0J (e 1) = d0 (e 1), ˙ − Σ( u, w ) − cle. Note that in uni-cyclic network models, W = { } (41) ∆J(f)F d0 = 0 either at f = 0 or if the motor is subject − The similarity between Eq. 33 and Eq. 41 is clear: both to an opposing stall force (f = Fs), where ∆J(Fs) = 0; of the expressions relate the velocity to the exponential thus the work production (W˙ ) is a non-monotonic func- of the free energy released upon the hydrolysis of ATP, tion of f, whereas E˙ and Q˙ decrease monotonically with and in both cases at equilibrium the motor does not F . move. However, the expression in Eq. 41 saturates at Although the conventional N-state unicyclic mod- large [ATP], in agreement with experiments (see Eq. 28). els (Fisher and Kolomeisky, 1999a, 2001; Kolomeisky and Compared to the case of a single chemical state, new Fisher, 2003) appear successful in describing the motil- load distribution factors are necessary for each transition ity and thermodynamics of molecular motors, unicyclic rate, i.e. models encounter two serious problems, especially when the molecular motor is stalled or starts taking backward −βθ+fd βθ−fd ui(f) = ui(0)e i , wi(f) = wi(0)e i . (42) steps at large hindering loads (Astumian and Bier, 1996; Hyeon et al., 2009; Liepelt and Lipowsky, 2007a). First, PN + the backward step in the unicyclic network, by construc- The distribution factors obey the relationship i=1(θi + − θi ) = 1 (Fisher and Kolomeisky, 1999b). It follows that, tion, is produced by a reversal of the forward cycle, which implies that the backward step is always realized via the N−1 Q + synthesis of ATP from ADP and Pi. This is the case i=0 ui(f) J (f) βX−βfd0 N−1 = − = e , (43) for rotary ATP-synthases, which function as transducers Q w (f) J (f) i=0 i of electrochemical potential into the synthesis of ATP in et al. and given ∆J(f) = J +(f) J −(f), the force-velocity the mitochondria (Alberts , 2008), and depending relation is, − on the sign of ∆µhyd and on environmental conditions (e.g. the presence and strength of a proton gradient QN and applied load), these spectacular motors can reverse wi(f)  v(f) = d ∆J(f) = d i=1 eβX−βfd0 1 . (44) their function (Itoh et al., 2004; Turina et al., 2003). Al- 0 0 Σ(f) − though ATP hydrolysis is reversible for linear molecular Although v(F ) is a complicated function of all the pa- motors (Bagshaw and Trentham, 1973; Hackney, 2005), rameters (Fisher and Kolomeisky, 1999b; Wagoner and back-stepping has not been associated with ATP synthe- Dill, 2016), the expression for a theoretical estimation of sis. Rather they are associated with ATP-independent the stall force is the same as the one found for N = 1, “slippage” or coupled to fuel consumption just as for- given by Eq. 37, and it is again equal to f in Eq. 10. ward stepping (Carter and Cross, 2005; Clancy et al., max et al. et al. In principle, one could design kinetic networks with an 2011; Gebhardt , 2006; Ikezaki , 2013). This arbitrary number of intermediates. However, because the requires an update of the kinetic network. number of free parameters becomes 4N 2, it is desir- In addition, the unicyclic network models lead to ˙ able to establish the minimal N that enables− an accu- Q = 0 under stall conditions (Eq. 46), which contradicts rate description of the experimental data as accurately physical reality. For example, an idling car still burns d 2 2 fuel and dissipates heat (Q˙ = 0)! There certainly exist as possible. Let D = limt→∞ dt ( x(t) x(t) ) be 6 d h i − h i fundamental differences between molecular world and the dispersion, and v = limt→∞ dt x is the velocity; the randomness parameter for a quantityh i x is defined as, macroscopic counterpart in that the former is subject to a large degree of fluctuations, which permit the reverse 2D process of ATP hydrolysis (i.e., synthesis) or negative r = . (45) heat dissipation for a sub-ensemble of the entire realiza- d0v tions. Yet, the mean of entropy production Q/T˙ is still It can be shown that N 1/r (Kolomeisky and Fisher, bound to be positive as demanded by thermodynamics, 2007; Koza, 2002), and therefore≥ r−1 sets a lower bound and the probability of a local violation of this principle to N that is needed to account for measurements. Note becomes vanishingly small as the system size grows. To that r is a function of ATP concentration and external ameliorate the aforementioned physically problematic load. interpretation, associated with the backward step 17

adjacent chemical states are given as

125D T 187 a hyd 126 k (f) = ko e−θfd0/kB T 188 127 [T] 25 25 189 128 o (1−θ)fd0/kB T 190 D T D k52(f) = k e 129 T 52 191 130 k (f) = 2ko (1 + eχij fd0/kB T )−1 for ij = 25 or 52192 (47) 131 ij ij 6 193 132 194 133 For the sake of simplicity, it is assumed that (2) 195(5), 134D T 196 b 125135 the step associated with the switching of the trailing197 and187 hyd 126136 leading head positions (the yellow arrow in Fig.8b),198 obeys188 137 199 127 [T] the Bell-like force dependence. When f > 0, the force189 D 138 200 D 128T T 190 129139 exerted on the motor hinders the forward step and201 en-191 140 202 130 [T] hances the backward step. Other steps (ij = 25 or 52)192 141 203 131 6 193 142 are abolished (kij 0) at large f(> 0), which204 corre- Analysis ofhyd kinesin-1 based onD the 6-state double cycle kinetic model. Schematics of the 6-state double cycle kinetic network for hand-over-hand dynamics of 132143 T Fig. 1. A. → 205 194 kinesin-1, where T , D, and „ denote ATP-, ADP-bound, and apo state, respectively.sponds Through binding to the [(1) (2)] physical and hydrolysis situation of ATP [(2) (5)], where and release the of ADP large [(5) (6)], external 133 æ æ æ 195 144 kinesin moves forward in the -cycle [(1) (2) (5) (6) (1)], where it takes a backstep in the -cycle [(4) (5) (2) (3) (4)]. The arrows in the 206 134 F æ æ æ æ force impedes theB chemicalæ æ stepsæ suchæ as ATP binding,196 145 figure depict the direction of reaction currents. In both cycles, each chemical step is reversible and the transition rate from the state (i) to (j) is given by kij . B. Reaction 207 135 current J , J , and J as a function of load. The three cartoons illustrate the amount of current along the and cycles at each value of load. f<0 and f>0 correspond 197 146 F B hydrolysis, andF ADPB release by deforming the confor-208 c to the assisting and hindering load, respectively. C. The ratio between the forward and backward fluxes J /J as function of f at fixed [ATP]. The stall forces, determined at 125 136147 F B 187 209 198 J /J =1(dashed line), are narrowly distributed between f =6 8 pN. (Dmation–H) V , D, Q˙ , ofW˙ , E molecular˙ as a function of f and motor. [ATP]. The whiteThe dashed model lines depict consists the locus of of two F B ≠ 126 137148 [ATP]-dependent stall force, and the dashed lines in magenta indicate the condition of [ATP] = 2 mM. I. Dependences of E˙ , Q˙ , W˙ on f at [ATP] = 2 mM. The188 star symbol in D 210 199 138149 indicates the cellular condition of f 1 pN and [ATP] 1 mM, which will be discussedsub-cycles later. ((1) (2) (5) (6) (1)) and211 200 127 ¥ ¥ F 189 B 128 139150 ((2) (3) (4) (5) (2)). At small190 (f 0)212 or as-201 140151 ≈ 213 202 129 motor eciency, if any, that is minimized. To evaluatesisting, hindering force load. (f < The0), backstep which in renders unicyclick reaction25191 schemes,k52, the reac- 141152 Q Q  214 203 Q˙ D V 130 142153 one should know , , and of the system (see Eq.2), whichtion currentby construction, is mainly is produced formed by a reversal along of the the192 forwardcounterclockwise cycle 215 204 Fig.can 1. beAnalysis gained of by kinesin-1 building based a proper on the 6-state kinetic double network cycle model kinetic model. that A.(15Schematics), which of implies the 6-state that double+ the cycle backstep kinetic network is always for hand-over-hand realized via dynamics 131 143154 direction of cycle ( ); in contrast,193 at large hindering216 205 of kinesin-1, where T , D, and „ denote ATP-, ADP-bound, and apo state, respectively. Through binding ((1) (2)) and hydrolysis of ATP ((2) (5)), and release of ADP 155 can delineate the dynamical characteristics of the system (11). the synthesisF of ATPæ fromF ADP and Pi. Moreæ importantly, 217 132 144 ((5) (6)), kinesin moves forward in the -cycle ((1) (2) (5) (6) (1)), where it takes˙ a backstep in the -cycle ((4) (5) 194(2) (3) (4)). The 206 156 Ouræ analysis on motors will showF that is aæ complexæ functionæ forceæ in (f calculating > 0) aboveQ from the kinetic stallB network, condition theæ unicyclicæ (f =æ reactionfs),æ the218 cur- 145ff depict the direction of reactionQ currents. In both cycles, each chemical step is reversible and the transition rate from the state (i) to (j) is given by kij . B. 207 133 157 ofstallf and [ATP], andstall that (f,[ATP]) is sensitive to a subtle scheme is led to Q˙ =0under the stall condition,195 which is 219 Reaction current J , J , and J asQ a function of load. The three cartoons illustraterent flow the amount along of current the established counterclockwise in and cycles at each direction value of load. ofC. Thecycle ratio 146 variation in motorF B structure. not compatible with the physicalF B reality that an idling car 208 134 158 between the forward and backward fluxes J /J as function of f at fixed [ATP].+ The stall forces, determined at J /J =1(dashed line), are196 narrowly distributedB between220 147 F B ( ),still which burns corresponds fuel and dissipatesF B to heatthe (ATPQ˙ =0 hydrolysis-induced). To build a 209 135 159 f =6 8 pN over the range of [ATP] = 10 µM 10 mM. (D–H) V , D, BQ˙ , W˙ , E˙ as a function of f and [ATP]. The white dashed line” in each197 panel depicts the221 locus of 148 ≠ ≠ backstep.more physically sensible model˙ that˙ ˙ considers the possibility of 210 136 FIG. 8 b. Schematics160 [ATP]-dependentResults of the and network stall Discussion force, and model the dashed representing lines in magenta indicate the the condition of [ATP] = 2 mM. I. Dependences of E, Q, W on f at [ATP]198 = 2 mM. 222 dynamics of double-headed149161 kinesin-1. T , D, and φ denote ATP-induced (fuel-burning) backstep, we extend the unicyclic 223 211 137 150 Owingreaction to scheme the microscopic into a multi-cyclic reversibility, scheme which199 takes each into cycle can212 ATP-, ADP-bound,162 Chemical and apo driving state, force, respectively. steady state current, Through and heat dis- 224 138 151163 networksipation modelof double that cycle can capture kinetic network the dynamical for kinesin-1. characteristicsThein principleaccountimportantly, an be ATP-consuming performed in calculating stall, inQ i.e.,˙ bothfrom a futile kinetic counterclockwise cycle200 network, (14–17). the225 uni- (+213 139 ATP binding152 [(1)164 →ofchemomechanical(2)], the system mechanical (11). dynamics Our step analysis[(2) of a→ onmolecular(5)], motors release motor will show can that be Incyclic+ the double+ reaction cycle scheme model, isled kinesin-1 to Q˙ =0 predominantlyunder201 the− stall moves− condition,226 214 sign,Q , ) and clockwise ( sign, , ) direc- of ADP [(5)→(6)],165 andmapped hydrolysis on an appropriate of ATP kinetic [(6)→ network(1)], kinesin where nonzero forward through -cycle under small hindering (f>0)or 227 140 153 is a complex function of f and [ATP], and that (f,[ATP]) FwhichB is not compatibleF with the− physical202 realityF thatB an idling 215 steady state current created by ATP hydrolysis freeQ energytions.assisting Forward load ( (backward)f<0), whereas steps it takes occur a backstep via˙ thethrough (2) (5) 141 moves forward154166 in theis sensitiveF-cycle to [(1) subtle→ (2) variations→ (5) in→ motor(6) → structure.(1)], car still burns fuel and dissipates heat203 (Q =0). To build228→ a 216 167 ( µhyd > 0) acts as the driving force. Given a proper net- the -cycle under large hindering load. In principle,” the 229 backward in the155 B-cycle≠ [(4) → (5) → (2) → (3) → (4)]. ((5) more(2))B physically transition; sensible the model edges that (1) considers(2) the ((2) possibility(1))217 142 156168 work model, the transport properties of the motor, V and current→ within the -cycle, J , itself is decomposed→204 into the →230 218 Fig. 1. Analysis of kinesin-1The based arrows on the in 6-state the figure double-cycle depict kinetic the model. directionA. Schematics of major of the coun- 6-state double-cycleand (4) kineticof ATP-induced network(5)+ ((5) for hand-over-handF (fuel-burning)(4))F are dynamics associated backstep, of itwith is indispensable ATP+ bind- 143 169 ResultsD, can be and expressed Discussion in terms of a set of rate constantsDRAFTkij , forward (J ) and backward current (J ≠), and205J = J 231 kinesin-1, where T , D, and „ denote ATP-,157 ADP-bound, and apo state, respectively. Through binding [(1) (2)] and{ hydrolysis} ofto ATP→ extend [(2)F (5)], the and→ unicyclic release of reaction ADP [(5) F scheme(6)], intoF a multi-cyclicF ≠ 219 terclockwise reaction currentk in each cycle. In both cycles,i j ≠ 170 Chemicalwhere ij is driving associated force, with steady the transition statecurrent, from theæ and-th to heating-th dis- (dissociation);J > 0 is satisfiedæ thein NESS. edges Although (6) æ a backstep(1) ((1) could be(6))232 and 144 kinesin moves forward in the -cycle [(1) 158(2) (5) (6) (1)], where it takes a backstep in the -cycle [(4) F(5)scheme(2) which(3) takes(4) into]. The account arrows an→ in the ATP-consuming206 → stall, i.e., 220 eachF chemicalæ step171 isstateæ reversible (4æ, 12, 13æ)(see andSIkijtext).defines The formal the transition expressionsB of V and ærealizedæ throughæ an ATPæ synthesis (18), corresponding to J ≠, 233 145 figure depict the direction of reaction currents.159 In bothsipation cycles, of each double chemical cycle step kinetic is reversible network and the for transition kinesin-1. rate(3) fromThe the(4) statea futile ((4)(i) to cycle(j)(3))is (14 given–17 are). by k associatedij . B. Reaction with207 ATP hydrolysisF 221 172 D in terms of kij (f,[ATP]) are used to fit simultaneously →experimental→ data (15, 19, 20) suggest that such backstep 234 current J , J , and J rateas a function from of the load.i160-th The three tochemomechanicalj cartoons-th state. illustrate{ c. the dynamicsReaction amount} of of current a molecular alongJF the, J motorBand, cycles can(synthesis) beat each value (The of load. cyanf<0 arrowsand f>0 incorrespondand cycles highlight222 146 F B 173 the experimental data of V and D obtained underF varyingB currentIn (ATP the synthesis 6-state double induced cycle backstep, model,J ≠ kinesin-1)208 is negligible predominantly in 235 to the assisting and hinderingand J load,as respectively. a function161 C.mapped ofThe load. ratio between on The an threeappropriate the forward cartoons and kinetic backward illustrate network fluxes J the where/J as nonzero function of f at fixed [ATP]. The+ stall forces, determinedF atF B 223 147 174 conditions of ATP and f (11). The set of rate constantsF B kATPij comparison hydrolysis).moves forward with J The through(ATP model hydrolysis-cycle in principleinduced under small209 backstep). accommodates hindering236 or as- J /J =1(dashed line),amount are narrowly of current distributed162 flowingsteady between state alongf =6 current the8 pN.F created(Dand–H)BV bycycles, D ATP, Q˙ , W as˙ hydrolysis, E a˙ as func- a function free{ of energyf} and [ATP]. The white dashedB lines depictF the locus of 224 F B 175 decided from the≠ fit further allows us to calculate the anity Thesisting dynamics load, whereasrealized in it the takes double a backstep cycle network through can the be 237-cycle. 148 [ATP]-dependent stall force, and the dashed lines in( magentaµ indicate> 0) acts the condition as the of driving [ATP] = force. 2 mM. I Given. Dependences a proper4 of E net-˙different, Q˙ , W˙ on f pathways:at [ATP] = 2 mM. The (i) star ATP-hydrolysis-induced symbol in D 210 B for- tion of f. f <1631760 and( ,f nethyd > driving0 are force), the and assisting reaction and current hindering (J) of the network, illuminatedIn principle, in terms the of current variation within of J and the J -cycle,with increasingJ , itself238 is de- 225 ≠A + F B 149 indicates the cellular condition of f 1 pN164 and [ATP]work1 model, mM, which the will transport be discussed properties later. of the˙ motor, V and + F 211 F 226 load, respectively.¥ 177 and¥ hence the heat dissipation from the motor, Q. wardf step(Fig.composed1B). ( Without); into (ii) the load ATP-hydrolysis-induced forward (f =0 (),J kinesin-1) and backward predominantly current backward239 (J ≠), 150 165178 D,For can kinesin-1, be expressed we considered in terms of a a6-state set of network rate constants (14), con-kij moves, + forwardF (J + J ). This imbalanceF diminishes212 with 240F− 227 step{ } ( and); (iii)J = forwardJF ∫ JB≠ step> 0 is that satisfied synthesizes at NESS. ATP Although ( a); 151 166179 wheresistingk ofij is two associated cycles, withand the(Fig. transition1A). fromThis theminimali-th to ki-j-thincreasingB Ff. AtF stall≠ conditions,F the two reaction213 currents 241B 228 F B (iv) backwardbackstep step could that be realized synthesizes through ATP an ATP ( synthesis−). The (18 re-) be obtained using a proper kinetic167180 networkstatenetic scheme(4, 12 model, 13 accommodates)(see thatSI cantext). 4 dialways Theerent formal kinetic realized expressions paths: via ATP- the of V synthesisare balanced of (J ATP= J from), so that ADP the and net current Pi. J associated 242 229 152 and a reversalF ofB the normal current direction214 F of cycle, i.e., 168181 andhydrolysisD in terms induced of forward/backwardkij (f,[ATP]) are step used and to ATP-synthesis fit experimentalactionwith currents the mechanical˙ JF steppingflowing defined through between (6) the states(1)F (2) and243 JB230 153 delineate the dynamicalmechanism, characteristics it has been of the proposed system (14{ to). OurextendMore} the unicyclic importantly, in calculatingJ ≠, experimentalQ from kinetic data ( network,15, 19, 20) the suggest215 that such backstep 169182 datainduced of V forward/backwardand D obtained step. under Although varying the conditions conventionalDRAFT of ATPand (5)F˙ vanishes (J = J J =0), but nonvanishing current 244 231 154 analyses on motorsnetwork show that intois a a networkcomplex function with multiple of f and cyclesunicyclic (Clancy network resultsthrough in currentQ (3)=0 (ATPunder(4) synthesis cantheF ≠ stall beB induced calculated condition, backstep,216 byJ ≠ the) is smaller generating than 183 (N=4)-stateand f simultaneously unicyclic kinetic (11). scheme The set confers of rate a similar constants resultkij due to chemistry still remains along the cycle ofF (2) 245 Q 170 J + (ATP hydrolysis induced backstep). æ æ 232 155 [ATP] and is sensitiveet al. to, 2011; a subtle Hyeon184 variationwithet the al.in 6-state motor, 2009; double structure. Liepelt cycle network andwhich Lipowsky, scheme is not at compatible small ffunction,it{ with}(3) the technique(4) physical(5) reality(6) by Koza(1) that an(2) (Hwang idling. A further and217 increaseHyeon,246 2017; 171 decided from the fit further allows us to calculate the anity æB æ æ æ æ æ 233 156 Transport and rotory motors studied185 is here ledare to a semi-optimized physically problematiccar interpretation still burns when fuel theand dissipatesof f beyondThe heatdynamics the stall (Q˙ force=0 realized renders). To in theJ build (J =0)orJ J224=0), but nonvan- 179 conventional (N=4)-state unicyclic kinetic scheme confersF a F ≠ B 241 163 chemomechanical3. dynamics Multi-cycle of Models180 a molecularsimilar result motor withcan the 6-state be doubleassisting cycle load network (f< scheme0),In whereas theishing double-cyclic it current takes duea backstep tonetwork chemistry through model still remains the225 total along heat the cycle gen-242 of (2) (3) (4) (5) (6) (1) (2) .A 164 mapped on an appropriate kinetic181 network,when f is which small, itallows is led tous a physicallythe -cycle problematic under interpre-erated large hindering fromæ the load.æ kineticæ In cycleprinciple,æ depictedæ the 226æ in Fig.æ 8bæ is de-243 B further increase of f beyond the stall force renders J 0 is satisfied in NESS.Q Although a backstep could be 229 subtleties in the185 physicsforward cycle of kinesin (15), which under implies externalF that the backstep load. is alwaysrent andcalculate affinity the (Barato rates of heat and dissipation Seifert, 2015; (Q˙ ), work Ge production and Qian,247 168 SI text). The formal expressions of V and D in terms of realized through an ATP synthesis˙ (21), corresponding to˙ J ≠, 230 In the model,186 the raterealized constants via the synthesis at the edges of ATP between from ADP the and Pi.2010; More Liepelt(W ), and and total Lipowsky, energy supply 2007a; (E)F forQian, varying 2004; conditions Qian and of 248 169 kij (f,[ATP]) are used to fit simultaneously theDRAFT experimen- experimental data (18, 22, 23) suggest that such backstep 231 { } 170 tal data of V and D obtained under varying conditions of ATP current (ATP synthesis induced backstep, J ≠) is negligible in 232 2 | www.pnas.org/cgi/doi/10.1073/pnas.XXXXXXX + F Hwang et al. 171 and f (14). The set of rate constants kij decided from the comparison with J (ATP hydrolysis induced backstep). 233 { } B fit further allows us to calculate the anity ( , net driving The dynamics realized in the double-cycle network can be 172 A 234 force), and reaction current (J) of the network, and hence the illuminated in terms of variation of J and J with increasing 173 F B 235 174 heat dissipation from the motor, Q˙ (see below). f (Fig.1B). Without load (f =0), kinesin-1 predominantly 236 175 For kinesin-1, we considered the 6-state network (17), con- moves forward (J J ). This imbalance diminishes with 237 F ∫ B 176 sisting of two cycles, and (Fig. 1A). This minimal ki- increasing f. At stall conditions, the two reaction currents 238 F B netic scheme accommodates 4 dierent kinetic paths: ATP- are balanced (J = J ), so that the net current J associated 177 F B 239 178 hydrolysis induced forward/backward step and ATP-synthesis with the mechanical stepping defined between the states (2) 240 179 induced forward/backward step. Although the conventional and (5) vanishes (J = J J =0), but nonvanishing current 241 F ≠ B 180 (N=4)-state unicyclic kinetic model confers a similar result due to chemistry still remains along the cycle of (2) 242 æ æ 181 with the 6-state double-cycle network model at small f (com- (3) (4) (5) (6) (1) (2) . A further increase 243 æ æ æ æ æ æ pare Figs.1 and Fig.S3), it is led to a physically problematic of f beyond the stall force renders J

2 | www.pnas.org/cgi/doi/10.1073/pnas.XXXXXXX Hwang et al. 18

however, because the chemical processes associated with ˙ βQ = β (JF AF + JBAB) “two” ATP hydrolysis events, one along the forward and     k12k25k56k61 k23k34k45k52 the other along the backward step, are still at work, the = JF log + JB log k21k52k65k16 k32k43k54k25 heat production is finite (Q˙ > 0)     k12k56k61 k23k34k45   = JF log JB log + (50) k12k23k34k45k56k61 k21k65k16 k32k43k54 βQ˙ = Js log = 2Js( ∆µhyd). | {z } k21k32k43k54k65k16 − chemical   (53) k25 + (JF − JB) log (51) k52 This is a point of great importance, which cannot be | {z } mechanical capture by the uni-cyclic network model (Fig.8a). Figure 8c depicts the currents flowing through the two sub-cycles JF and JB calculated from the set of rate con- stants ( k (f) ) determined against the motility data Beard, 2005; Seifert, 2012) { ij } of kinesin-1. Under small hindering (f & 0) or assist- Q˙ = JF F + JB B, (48) ing load (f < 0), kinesin-1 predominantly moves forward A A through the -cycle, whereas it starts taking more num- where the two driving forces (affinities) are given by, F + A ber of backsteps through the pathway as the load   increases further. At stall conditions, theB net current as- k12k25k56k61 (a) F = kBT log = ∆µhyd fd0 sociated with the mechanical stepping defined between A k21k52k65k16 − −   the states (2) and (5) vanishes (J = JF JB = 0); k23k34k45k52 − (b) B = kBT log = ∆µhyd + fd0 however, non-vanishing current along the futile cycle hy- A k32k43k54k25 − drolyzing ATP still persists along the reaction path of (49) (1) (2) (3) (4) (5) (6) (1) (see Figure Note that at f = 0, the chemical driving forces for F 8b). A further increase of f beyond fs leads to JF < JB, and cycles are identical to be ∆µhyd. The above de- increasing the chance of backsteps via ATP hydrolysis. compositionB of affinity associated− with each cycle into Although a backward step satisfying JF < 0 could be the chemical driving force and the work done by the realized through an ATP synthesis, a theoretical analy- motor follows naturally from the f-dependent expression sis (Hyeon et al., 2009) of the experimental data (Carter of kij given in Eq.47 (Fisher and Kolomeisky, 1999a, and Cross, 2005; Nishiyama et al., 2002) suggest that 2001;{ Liepelt} and Lipowsky, 2007a). The entropy pro- such event (JF < 0, ATP synthesis induced backstep) is duction from the double cycle model is expressed as the practically negligible compared with the one associated sum of entropy produced from the two sub-cycles and with ATP hydrolysis-induced backstep (JB > 0), so that . Or simply from Eq. 48 and Eq. 49, Q˙ canF be re- B JB JF . This is in agreement with energetic analysis cast into the differnece between the total free energy suggesting| |  | | that backward and forward steps are not the input [E˙ = (JF + JB)( ∆µ )] and work production − hyd reverse of each other (Hackney, 2005). [W˙ = (JF JB)fd ]: − 0 Q˙ = JF F + JB B A A D. Additional remark on non-equilibrium nature of motors = (JF + JB)( ∆µ ) (JF JB)fd − hyd − − 0 = E˙ W.˙ (52) It is worth mentioning a couple of recent studies that − further highlight the non-equilibrium aspect of molecular Notice that in order to reiterate our earlier remark that motors. First, the Harada-Sasa equality (Harada and the mechanical equilibrium at the stall condition does not Sasa, 2006) quantifies the heat generated from Langevin correspond to the thermodynamic equilibrium, the last processes out of equilibrium as follows, line of βQ˙ (Eq.51) is decomposed into the contributions N ∞ by chemical and mechanical processes. In the proposed X  Z dω  Q˙ = γ v2 + γ [C˜ (ω) 2k T R˜0 (ω)] . double-cycle kinetic model, the average velocity of the i i ii − B ii 2π i=1 −∞ motor is given by V = d0(JF JB). Thus, if the num- (54) bers of forward and backward− steps taken over time are balanced to each other, satisfying JF = JB (Carter and It equates the total heat dissipation rate from the sys- Cross, 2005), we expect no net directional movement of tem Q˙ with the expression in terms of friction coefficient the motor ( x(t) = 0), which is equivalent to say V = 0. γi of an i-th variable xi, the Fourier component of auto- It is importanth toi recognize that displacement (or travel correlation of the velocity fluctuation C (t) (x ˙ (t) ii ≡ h i − distance) x(t) is the most direct observable to an exter- vi)(x ˙(0) vi) 0, and the real part of the response func- ˜0 − i nal observer if one were to use optical tweezers or fluores- tion Rii(ω). At equilibrium, when the detailed balance cence dyes. In the stall condition (JF = JB J = 0), condition is satisfied, v = 0, which leads to the standard ≡ s 6 i 19

˜ ˜0 fluctuation dissipation theorem, Cii(ω) = 2kBT Rii(ω). E. Applications Thus, the heat dissipation is zero. Therefore, the equality relates the extent of violation of the fluctuation-response 1. Myosin V and Kinesin-1 relation in NESS with the rate of energy dissipation from the system into the bath. At least two studies have so SKMs have been successfully used to model the func- far used this equality to experimentally assess the heat tion of molecular motors. Sometime ago Leibler and dissipation from molecular motors, one for kinesin and Huse (Leibler and Huse, 1993) devised a model to investi- the other for F1-ATPase. Ariga (Ariga et al., 2018) have gate the differences between motors that act as “porters” adapted this equality in the form, and those that function as “rowers.” The former, like  Z ∞  the processive molecular motors described so far, work ˙ 2 ˜ ˜0 dω in small groups in order to transport cargoes; myosin Qx = γx vx + [Cvv(ω) 2kBT Rvv(ω)] (55) −∞ − 2π II and axonemal dynein are examples of the latter, and work in large teams in order to slide or bend filaments. where the displacement of kinesin motor x(t) was experi- In two landmark papers Fisher and Kolomeisky an- mentally monitored, and the Fourier component of auto- alyzed the data for kinesin-1 (Fisher and Kolomeisky, correlation function and response function of velocity 2001) and myosin V (Kolomeisky and Fisher, 2003) in fluctuation were directly calculated using the time traces order to train the parameters in their SKMs. A 2-state obtained from single molecule measurements. They and a 4-state model for kinesin-1 described accurately showed that the total heat dissipation assessed using the motor velocity as a function of ATP and resistive Eq.55 and the work done by the kinesin do not add up to load (see Fig. 9a-b); although the 2-state model was also the total chemical free energy input to the motor. The used to describe the run-length of kinesin (see the sec- authors suggested that there are other elements of heat tion VI for details about the run length), the analysis of dissipation that do not involve the dynamics of the motor the randomness parameter as a function of [ATP] and along the direction of motion x(t) that can be monitored f highlighted the need for a N = 4 model. In the case by their experiment. Although the proposed double cycle of myosin V, Kolomeisky and Fisher considered the data model in Fig.8b has already accommodated other possi- for ATP and load-dependent dwell time before a forward bilities, such as ATP-hydrolysis involved futile cycle with- step (Mehta et al., 1999); they fit the parameters of their out stepping, other scenarios such as slippage induced by 2-state model (Fig. 9c) and predicted the value of the - mechanical force without chemical process (Yildiz et al., domness parameter as a function of [ATP] and f. The 2008) are not included. Further elements of heat dissi- analysis of the parameters trained with the experimen- pation could be included in kinetic models in order to tal data revealed in both circumstances the existence of describe the function of the motors. a substep, and that most of the load dependence on the The finding by Ariga et al. should be contrasted rates was carried by the reverse processes, indicating that with another experimental study using the Harada-Sasa the transition states are closer to the initial state than to equality on a single F -ATPase (Toyabe et al., 2010). 1 the final state. These authors showed that the total free energy input to the enzyme was partitioned to the heat and work Liepelt and Lipowsky (Liepelt and Lipowsky, 2007a) production with little loss. This indicates that the cycle devised a multi-cycle model (see Fig. 8b), which allowed of ATP synthesis corresponds to the reversed cycle of them to incorporate backward steps fueled by ATP hy- the hydrolysis-driven motor rotation. However, a recent drolysis. The experimental results of Carter and Cross careful theoretical analysis by Sumi and Klumpp on (2005) and Visscher et al. (1999) were recovered with a di-cyclic, N = 6 model (see Fig. 10), although Liepelt F1-ATPase (Sumi and Klumpp, 2019) suggested that the reversibility (100% efficiency) of the rotary motor and Lipowsky showed that another state (and a new cy- is only attained under certain conditions. Mechanical cle) was necessary in order to account for the velocity as slip can occur in the presence of high external torque a function of ADP concentration (Schief et al., 2004). without chemomechanical coupling, the effect of which A further development was put forth by Hyeon et al. is amplified at low ATP and ADP concentrations. This (2009), where a more sophisticated network was proposed reduces the previously estimated 100 % free-energy in order to consider an alternative pathway for backward transduction efficiency. It was argued that in addition stepping based upon known structural features of the to the viscous dissipation of the probe, heat dissipa- kinesin-1 dimer. Clancy et al. (2011) proposed yet an- tion could occur from the rotary motor itself as the other cycle for kinesin, in this case to fit the data for a torque applied to the biological nanomachine induced mutant with an extended neck linker; the model success- mechanical slip. This occurs as a result of deformation fully recovered the velocity, randomness parameter, and of molecular conformation, which is best optimized to ratio between number of forward and backward steps as a function in the absence of torque. function of ATP and resistive load, even in the super-stall regime. 20

FIG. 10 Fit of experimental data for kinesin-1 with a di-cycle, N = 6 model. The data in panels (a) and (b) is from Carter and Cross (2005), whereas the experiments considered in pan- els (c) and (d) are from Visscher et al. (1999). The figure is reproduced from Liepelt and Lipowsky (2007a).

Fig. 3A in (Tsygankov et al., 2011)). A much more elabo- rate model that couples structural aspects of dynein with a model for the catalytic cycle, which is similar in spirit FIG. 9 Unicycle models for kinesin-1 and myosin V. Panel to the theory for myosin V (Hinczewski et al., 2013), (a-b): comparison for kinesin-1 of the results of a uni-cycle was proposed more recently (Sarlah and Vilfan, 2014). model with experiments from Visscher et al. (1999). Panel (c): myosin V, comparison between a N = 2 model for The predictions of the model, which has a large number myosin V and the velocity from experiments, obtained from of parameters, were successful in obtaining the step-size the mean dwell times τ (Mehta et al., 1999) via the rela- distribution as well as an estimate of the stall force. How- tionship v = 36nm/τ. Panels (a-b) and panel (c) are adapted ever, simple theories that can account for the unusually from Fisher and Kolomeisky (2001) and from Kolomeisky and broad step size distributions reflecting the erratic nature Fisher (2003), respectively. of this motor, force-velocity curves as a function of ATP concentration, as has been done for kinesin and myosin V, are lacking. 2. Dynein – an erratic motor

In part due to its complex architecture and the paucity VI. MOLECULAR MOTORS – MODELS WITH of details of the nucleotide chemistry much less is known DETACHMENT about the kinematics of dynein stepping kinetics. We should note that both from the perspective of structure All molecular motors take only a finite number of steps determination (Schmidt et al., 2015) and single molecule along their tracks before detaching. Therefore, promi- experiments (Belyy et al., 2014b; DeWitt et al., 2012; nent features of motor motility are the run-length, L, Reck-Peterson et al., 2006a) there have been spectacular and the run-time, T , which correspond to the distance advances in recent years, which could be used to create covered during a processive run, and the amount of time new theories. A theoretical model different from those spent bound to the filament before detaching, respec- described in this perspective, which by necessity treated tively. In addition, the ratio between L and T consti- the structural and transition kinetics between pre- and tutes an alternative and intuitive definition of velocity, post-power stroke states approximately, was introduced v = L/T , which is simply the ratio between the spa- (Tsygankov et al., 2011) to calculate the distribution of tial displacement of a motor and the amount of time step size. Despite several untested approximation, exper- taken to complete the movement. Processive motors take iments (Reck-Peterson et al., 2006a) and kinetic Monte many steps along the filament before dissociating from Carlo simulations of the model were in fair agreement (see the track. The number of steps depends on the motor, 21 and on a number of parameters such as the concentration �⃗

of nucleotides, and the value of the applied load. This � A B C � underscores the importance of accounting for the end of

the processive run in a way that is consistent with the en- � - +

k k zymatic cycle of a specific motor and the experimentally �

Activation Energy �

determined run-length, run-time etc. i-1 i i+1 |� | |� | The models described in the previous sections assume that the motor is permanently attached to the track,

3pN 0.6 4pN A B 0.4 D E and ignore the end of the processive run. We discuss 0.0030 5pN

P(v=0) 0.0006 7.6pN 0.2 8.5pN 0.0 0 2 4 6 8 here a number of strategies that have been used in 0.0020 F (pN) order to account for the detachment from the filament. P(v) 0.0003 (i) The most direct way is to increase the number P(v≠0) 0.0010 0.0000 of states by Nd, the number of detached states, and 0.0000 -400 -200 0 200 400 200 100 0 100 200 to introduce the rates of releasing and attaching to v (nm/s) v (nm/s) the filament. This method, which was used to model the function of kinesin-1 with Nd = 1 (Liepelt and Lipowsky, 2007a), enables a steady-state treatment of FIG. 11 One-state model for kinesin motility with detach- the corresponding Markov jump process, but requires an ment. (A) Kinesin walks along the MT, subject to an external increase in the number of states, the number of edges, load F . (B) The MT is represented as a linear discrete set of sites separated by d0 = 8.2nm; the motor advances towards and the number of cycles. (ii) Alternatively, it is possible the + end with rate k+, backsteps with rate k−, and at ev- to enforce a new stationary state in the Markov jump ery site γ establishes the detachment rate. (C) Load affects process by diverting the flux into the detached state the rates according to Bell model, which in turn corresponds towards the filament-bound conformations. Although to a modification of the energy profile: k+ diminishes as load this approach does not require any other state (N = 0), increases, whereas the resistive force favors back-stepping (in- d − it is necessary to increase the number of edges, which creases k ). The transition state location with respect to the |d−| results in the creation of new cycles. A theoretical starting state, θ = (|d−|+|d+|) . (D) The velocity distribution description of this steady-state approach can be found displays a markedly bimodal shape at a variety of resistive in the remarkable work by Hill (Hill, 2005a). It is forces, with the negative peak that increases as the load ap- unlikely that the methods described by Hill could be proaches stall. (E) Simulations performed assuming an exper- imentally motivated error on the determination of d0 maintain used to calculate velocity and run-length distribution. the bimodal structure of P (v). Figure adapted from Vu et al. However, one could calculate their averages as a function (2016). of nucleotide concentration or external load. (iii) It is also possible to account for the detached trajecto- ries by re-normalizing quantities such as the average A. Velocity Distribution velocity over the decreasing number of bound motors. This strategy, put forth by Kolomeisky and Fisher, is The simplest model of molecular motor with detach- described in (Kolomeisky and Fisher, 2000), and has ment is shown in Fig. 11, in which forward (backward) been used to test an approximate equation for kinesin steps occur with rate k+ (k−), and the detachment rate run length (Fisher and Kolomeisky, 2001). (iv) One is γ. For this model it is possible to determine analyti- could define the average run length as the ratio between cally the probability distribution p(¯v), where the veloc- the average velocity v and the rate of detachment γ ityv ¯ is the ratio between the net number of steps taken (L v/γ). In order to do so, γ could be estimated (n = m l, where m is the number of forward steps and ≈ as the product between the stationary probabilities of l is the− number of backward steps) divided by the run- occupying the “vulnerable” states and newly introduced time (¯v = n/T = v/d0). Vu et al. showed that (Vu et al., P S detachment rates (γ = i∈vulnerable pi ki,det) (Fisher 2016), and Kolomeisky, 2001; Maes and Van Wieren, 2003). (v) Finally, the detached state could be treated as an p(¯v ≷ 0) = ∞ absorbing state. A stationary solution is not possible: γ  n n+1 1 n2k+k− −1 X ± −kT/|v¯| n after a sufficient time (t >> γ ) all the motors will = (k e ) 0F1(; n + 1; ), v¯ v¯ n! v¯ 2 be absorbed to the detached state. However, one may | | n=0 | | | | account for all of the possible trajectories leading to (56) n2 + − absorption, and perform averages over the ensemble of where 0F1(; n+1; |v¯|2 k k ) is a hypergeometric function, + − these paths. In the remainder of this section we focus and kT = k + k + γ. For an alternative derivation, see on this last method. a more recent study (Zhang and Kolomeisky, 2018). This model was adopted to describe the stepping mechanism 22 of kinesin-1; the parameters k+ and k− were established from experiments and the run length and velocity distri- detached state bution (Walter et al., 2012) were fitted using only one pa- rameter, γ. The result in Eq. 56 hold also in the presence γ

of force, which increases the probability of back-stepping → and detaching from the microtubule. Using this model a → number of noteworthy results were obtained (Vu et al., 2016): (i) even at zero load, the velocity distribution is k− 2l k+ not Gaussian, although it approaches a normal distribu- → tion if the run length is large (that is, k+/γ >> 1 and k → k+/k− >> 1); (ii) the probability distribution of the in- stantaneous velocity (the size of the step divided by the dwell time) differs from Eq. 56, and does not match the experimental distribution (Walter et al., 2012), suggest- ⋯1l−1 1l 1l+1⋯ ing that it is not the correct way to compute velocity; (iii) FIG. 12 Two-state model with detachment. The forward and most surprisingly, the velocity distribution is bimodal, backward stepping rates are k+ and k−, respectively, and they with distinct peaks forv ¯ > 0 andv ¯ < 0. The prediction connect state 2 to state 1. The 1 → 2 transition occurs with that p(v) is bimodal, becoming most prominent at stall rate k. Detachment occurs only from state 2, with rate γ. force, was unexpected. At f = fstall, the na¨ıve expecta- tion would be that p(v) would have a peak at v = 0, and head is still bound to the MT (or at least in the vicinity the area under p(v) with v > 0 and v < 0 would be the of the pre-stepping site) (Mickolajczyk et al., 2015), same. The bimodal distribution is a consequence of the the other group suggested that the dissociation from discrete nature of kinesin step size, and it is exaggerated the MT and forward movement of the trailing head at large f, when the probability of taking backward steps detachment precede ATP binding (Isojima et al., 2016). and the probability of detaching are larger. Using the model in Fig. 12, Takaki et al. (Takaki et al., The model in Fig. 11 was recently extended in order 2019) predicted that the ATP-dependent profile of to include an intermediate step (Fig. 12), which was used the randomness parameter may be used in order to to describe the kinetics of a molecular motor as a com- determine whether ATP binds to the two-head-bound or bination of an ATP-dependent and an ATP-independent one-head-bound conformation of the kinesin-1 dimer. transition (Takaki et al., 2019). The resulting velocity distribution is, p(¯v ≷ 0) = B. Alternate Models with Detachment γ X m l √π kn+1(k+)m(k−)l m l n+1/2 − − v¯ v¯ m!l! k (k+ + k− + γ) n+1/2 v¯ m,l | − | More complicated kinetic schemes that include mo- m≷l tor detachment have been used in order to compute the + − + − − k+k +k +γ m−l  k (k + k + γ) m l  processivity and velocity of molecular motors. Elting et e 2 v¯ I | − | − , n+1/2 2 v¯ al. (Elting et al., 2011) found an analytical solution for (57) a 8-state model of myosin VI which they proposed in or- where I is a modified Bessel function of the first kind. In- der to investigate the nature of the gating mechanism. terestingly, the bimodal structure of the velocity distribu- This model did not account for the possibility of back- tion is robust to changes of the nucleotide concentration, ward stepping, which was instead included by Caporizzo and it is independent of which step (1 2 or 2 1) et al. (Caporizzo et al., 2018) in a kinetic model that was → → is ATP-dependent. Therefore, an experiment aimed at devised in order to investigate the motility of myosin X testing the predicted multi-modality of p(v) may be con- as a function of the structure of the filament (individ- ducted under arbitrary ATP concentration and resisting ual versus bundled actin) and the geometry of the tail of load. the dimer (parallel versus anti-parallel). More recently, a The model in Fig. 12 was used to tackle a vexing new theoretical framework capable of extracting motility conundrum concerning the stepping mechanism of characteristics such as average velocity, average and dis- kinesin (Takaki et al., 2019). Recently, two groups have tribution of number of steps, and probability of backward performed similar experiments in which they monitored stepping for a kinetic network of arbitrary geometry has the movement of one kinesin head conjugated with been proposed (Mugnai et al., 2019). The dynamics of a gold-nanoparticle (AuNP) by tracking the position the motors is described in terms of the following Markov of AuNP (Isojima et al., 2016; Mickolajczyk et al., chain, 2015). From the analysis of the trajectories, one group proposed that ATP binds to kinesin when the trailing P~x+1 = (Sˆ + Fˆ + B)ˆ P~x = Mˆ P~x, (58) 23

tation (it only requires matrix algebra). Therefore, it detachment can be used to fit complicated models against experi- mental data. The model was used to solve a compli- cated model for myosin VI motility inspired by model of Yanagida and coworkers (Ikezaki et al., 2012, 2013; B̂ F̂ Nishikawa et al., 2010), which accounted for a motor tak- ing both hand-over-hand and inchworm-like steps (Mug- nai et al., 2019), and predicted the gating mechanism [in Ŝ agreement with (Dunn et al., 2010; Elting et al., 2011)], size of the backward steps [matching experimental ob- l − 1 l l + 1 servations (Altman et al., 2004; Nishikawa et al., 2010) FIG. 13 Model for processive motor moving on a periodic within discrepancies likely related to fluctuations and the track. From each site l of a filament the motor can change site of probe attachment], pathway for backward stepping its conformation (e.g. the chemical state) within the same [as suggested by (Ikezaki et al., 2013)], and importance location (transitions included in the matrix Sˆ, in black) move of foot-stomping in breaking the tight-coupling of myosin forward (the matrix Fˆ, in red, accounts for these steps) or VI stepping. backward (the matrix Bˆ, in blue, refers to these displace- ments) along the track, or detach (magenta). VII. POLYMER PHYSICS-BASED APPROACHES ~ INCORPORATING STRUCTURAL FEATURES INTO where the N-dimensional vector Px contains the proba- KINETIC THEORIES bility of occupying any of the N filament-bound states of the motor after x transitions have occurred, and S,ˆ Even with coarse-grained numerical models like the one F,ˆ and Bˆ are the N N transition probability matrices proposed by Craig and Linke (2009),which is described for undergoing a transition× at a fixed location, stepping below, collecting statistics from multiple simulations cov- forward, or backward, respectively (see Fig. 13). Eq. 58 ering entire motor trajectories (i.e. for myosin V each run holds under the assumption that the track is periodic, averaging tens of steps until detachment) can be compu- which makes S,ˆ F,ˆ and Bˆ independent on the location of tationally expensive. To more comprehensively explore the motor along the filament. By accounting for all of the how motor architecture affects stepping dynamics, par- possible stepping pathways, it was found that the prob- ticularly in light of experiments that perturb structural ability of taking n steps (forward or backward) before features like lever arm length (Oke et al., 2010; Sakamoto detaching is, et al., 2005), alternative approaches are needed. We fo- | ˆ ˆ ˆ n ~ cus on one particular example, an analytical theory for P(n) = ~1 (I Pstep)(Pstep) P0, (59) − myosin V dynamics (Hinczewski et al., 2013), that also where ~1| is the transpose of an N-dimensional vector of highlights several aspects discussed earlier: the starting ones, P~0 is the initial probability (appropriately normal- point of the theory is a six-state stochastic kinetic net- −1 ized, ~1| P~0 = 1), and Pˆstep = (Fˆ + B)(ˆ ˆI S)ˆ . This work model consisting of multiple cycles that explicitly expression· is a generalization of a well-known− result for includes detachment of the myosin V dimer from the N = 1, in which P(n) = (1 π)πn, where π < 1 is actin filament. Incorporating detachment allows us to the probability of stepping and− 1 π is the probability model finite run lengths of the motor on actin, while of detaching. The average number− of steps n may be the multi-cycle network topology captures the dominant computed using simple linear algebra, as wellh asi the dis- pathways of myosin V dynamics under load forces below tribution and average number of forward and backward or near stall ( 2 3 pN): these include both forward . − steps, m and l , respectively, with m + l = n . and backward steps as well as so-called “stomps”, where h i h i h i h i h i The average run length is then L = d0( m l ), either the trailing or leading head detaches and then reat- and by using a classical result for theh i mean-first-passage-h i − h i taches near its original binding location. Stomps are time to absorption it is possible to compute τ (Op- challenging to detect with single-molecule fluorescence penheim et al., 1977), and define an averageh velocityi techniques, but have been observed experimentally us- v = L / τ . With Kinetic Monte Carlo simulations it is ing high-speed atomic force microscopy (Kodera et al., possibleh i toh i show that L / τ differs from L/τ , which 2010). was adopted in (Takakih eti al.h ,i 2019; Vu et al.h , 2016).i On However unlike the kinetic models treated so far, the other hand, it was shown empirically that the dif- the force-sensitive transition rates will not be described ferences become smaller as the average number of steps through a phenomenological Bell-like exponential depen- taken by the motor increases (Mugnai et al., 2019). The dence as in Eq. (35). Instead, the goal of the approach advantages of this theoretical framework are its flexibil- by Hinczewski et al. (2013) is to model the structural ity (it works with any network) and ease of implemen- mechanics underlying this force dependence through a 24 coarse-grained polymer theory for the myosin V dimer. rize the underlying six-state kinetic model, illustrated in This replaces the discrete semiflexible polymer descrip- Fig. 15A. State 1 is the waiting state, where both mo- tion of Craig and Linke (Craig and Linke, 2009) (inter- tor heads have bound ADP and are strongly attached acting monomers representing the motor head and lever to actin. The lever arms for both heads are in the arm IQ domains) with an even simpler construct: a sin- post-powerstroke conformation: in other words in the gle continuous semiflexible polymer that represents the absence of other constraints the lever arms would ori- combined motor head and lever arm domains. Thus, the ent in the forward direction (toward the plus end of dimer becomes two polymer “legs” attached at a flexible actin), but because the junction pulls the lever arm of junction. The load force transmitted through the cargo the leading head backwards the entire structure is in a domain is modeled by an effective force F applied at the strained state known as the telemark or reverse arrow- junction. Because the load force changes the ensemble head stance (Kodera et al., 2010; Walker et al., 2000). of conformations for the two polymer legs, and hence State 10 is also a waiting state, but with the entire mo- the three-dimensional distribution of positions explored tor displaced forward by one half helical length of actin by the unbound motor head (see Fig. 14), it affects the (∆ = 36 nm). The possible kinetic pathways are as fol- rates at which the unbound head reaches potential bind- lows: ing sites. 1. Forward step (1 2 3 4 10): ADP release The most direct advantage of this simpler coarse- → → → → grained description is that the effects of force on stepping from the trailing head is followed by ATP binding, dynamics become (to an excellent approximation) ana- which leads to detachment of the head from actin (1 2). We assume saturating ATP conditions, lytically tractable. It is also readily generalized to more → complex contexts. For example, while the model we fo- so ATP binding is fast compared to ADP release, cus on here considers only a flexible junction, backwards and hence the entire process is subsumed into a sin- −1 load forces parallel to the actin axis, and restricts steps to gle transition described by a rate td1 . Here td1 is actin binding sites separated at actin helical half lengths the mean detachment time scale, dominated by the −1 (the most probable step separation for myosin V), all of waiting time for ADP release, td1 12 s (De la Cruz et al., 1999). The next transition∼ (2 3) in- these assumptions can be relaxed. A recent extension of → the polymer theory approach (Hathcock et al., 2019) ex- volves ATP hydrolysis into ADP+Pi, along with a plores the effects of a possible structural constraint at the recovery stroke where the head / lever-arm orien- junction (inspired by experimental evidence (Andrecka tation changes from post- to pre-powerstroke. This −1 −1 et al., 2015)), considers off-axis forces, and incorporates transition occurs at a rate th = 750 s (De la the full distribution of steps at all possible actin binding Cruz et al., 1999). While there are scenarios where sites. This allows us to understand previously observed the reverse hydrolysis rate is significant (for exam- step distributions of mutant myosins with various lever ple in motors with modified light chain composi- arm lengths (Oke et al., 2010; Sakamoto et al., 2005), as tion (Dunn and Spudich, 2007)), for this discussion well as the robustness of myosin V dynamics in experi- we make the typical assumption that the forward ments where various off-axis load forces are applied to the hydrolysis rate dominates. In general, while ev- motor via optically trapped cargo (Oguchi et al., 2010). ery transition arrow in Fig. 15 has an associated Moreover the polymer approach is not limited to myosin reverse transition in principle, here we make the V: an analogous treatment of dynein, with the two motor simplifying assumption that the reverse rates are domains approximated as rigid rods (the large stiffness negligible relative to the forward rates. This as- limit of semiflexible polymers), can successfully repro- sumption captures the dominant kinetic pathways duce the complex details of dynein step distributions on of the motor (our focus here), but we would need to microtubules (Goldtzvik et al., 2019). For both myosin V explicitly consider reverse rates to look at thermo- and dynein each half of the dimer is modeled as a single dynamic features of the system (like entropy pro- polymer / rod, and this description is likely to be appli- duction). Once ATP is hydrolyzed, the motor head cable to other dimeric processive motors with fairly stiff has the ability to associate with actin again. The lever arms, like myosin XI (Tominaga and Nakano, 2012). three-dimensional diffusive search can result in ei- However there are cases, like myosin VI or X, where ther binding to the forward site (36 nm ahead of the the lever arm structure is more heterogeneous, mixing bound leg), or the original binding site. The rate at + −1 stiff and flexible domains (Sun and Goldman, 2011b). which it binds to the forward site is (tfp) , which + Here any future attempt to apply coarse-grained poly- depends on the mean first passage time tfp to the + mer modeling would need to represent each lever arm as forward site. The dependence of tfp on the load a series of polymer chains with different bending rigidi- force is related to how the force biases the diffusive ties. search, and the underlying physics is discussed in To understand the basic details of the polymer the- more detail below. A binding to the forward site ory for myosin V in its simplest form, we first summa- (3 4) is quickly followed by P release and a pow- → i 25

FIG. 14 Examples of the diffusive search trajectory of the detached myosin V head for two different values of backwards load force: A. F = 0; B F = 2 pN. The actin filament lies along the z axis, and the trajectories are projected onto the radial distance px2 + y2 away from the actin axis. The colors represent time, with lighter colors (yellow) occurring earlier than darker colors (red). In panel A the trajectory ends in a forward step (the detached head going from the backward binding site at z = −36 nm to the forward binding site at z = 36 nm. In contrast, panel B shows a backwards step, which becomes increasingly favored as a kinetic pathway under larger load forces. Superimposed on the trajectories are cartoon snapshots of the two polymer legs (each leg representing a lever arm plus head) of the coarse-grained model for myosin V, at a time shortly after detachment. The white dot is the junction of the two legs, and the cyan dot is the location of the detached head at that time step. The contours correspond to the equilibrium probability distribution P(r) of the detached head, calculated from the analytical polymer theory (darker contours correspond to higher probabilities). As load force is increased, the attached leg and junction are pulled backwards, biasing the entire distribution away from the forward binding site, and thus increasing the chance of backwards versus forward steps (see also the corresponding experimental data of Fig. 16A). Adapted from Hinczewski et al. (2013).

A termination B 100 forward step ATP ADP P 2 i 3

ADP ADP leading stomp 10−1 1 1' 5 trailing stomp forward step 4 termination −2 ADPADP ADP Pi ADP ADP ADP Pi ADPADP 10

leading stomp backward step 6 Pathway probability backward 10−3 step

ADP ADP trailing stomp

10−4 0.0 0.5 1.0 1.5 2.0 termination Force F [pN]

FIG. 15 A. Six-state kinetic model for myosin V that underlies the coarse-grained polymer analytical theory of Hinczewski et al. (2013). States 1 and 10 are identical waiting states, with both motor heads bound to ADP and actin, except that 10 is displaced 36 nm (a half helical length of actin) toward the plus end of the filament. Arrows are marked by inverse timescales that denote the transition rates, the details of which are described in the text. Colored arrows are transitions that belong to specific kinetic pathways: forward step (dark blue); trailing stomp (green); leading stomp (purple); and backward step (red); termination (light blue). Gray arrows are transitions that are shared between multiple pathways. B. Predicted pathway ± probabilities from the model, using the kinetic network together with mean first passage-times for the diffusive search step tfp from polymer theory. Adapted from Hinczewski et al. (2013). 26

0 −1 erstroke (4 1 ), which occurs with a rate tps . We arm has to adopt a strained stance) and so as be- assume this→ step is rapid relative to the others in fore we introduce a factor b to scale the rate to the + the network, such that tps td1, tfp, th, and hence forward site. Thus the overall reattachment rate  0 + −1 tps will not enter explicitly into our estimates for from 6 1 is b(tfp) . motor properties below. → 4. Backward step (10 6 1): This pathway starts → → 2. Trailing stomp (1 2 3 5 1): This path- with leading head detachment (10 6) like the → way starts in the→ same→ way→ as the→ forward step, leading stomp, but the detached head after diffu- with unbinding of the trailing head (1 2) fol- sion finds the backward binding site. The post- lowed by hydrolysis and the recovery stroke→ (2 powerstroke conformation with ADP bound is ge- 3), but the diffusive search now ends at the orig-→ ometrically favorable for binding to this site (be- inal binding site from where the head detached cause the lever arm does not have to be bent) and − −1 (3 5). The mean first passage time to this site hence the rate from 6 1 is (tfp) . Note that this →− → is tfp (which also depends on force as described description of backward stepping is approximately below). However the head no longer binds in the valid for load forces near or below stall, but does same configuration as before, because the lever arm not include additional features that may become has undergone a recovery stroke and is now in the important for larger superstall forces (F > 3 pN), pre-powerstroke orientation. Thus binding requires such as power stroke reversal (Sellers and Veigel, not only first passage to the site but overcoming the 2010). energetic barrier of an unfavorable geometric orien- 5. Termination: For the kinetic pathways described tation. We effectively model this through a factor above, if the motor is in one of the three states 0 < b < 1 that reduces the rate, so that the overall where only one head is bound to actin (2, 3, or rate from 3 5 is given by b(t− )−1. From fitting fp 6), then detachment of the second head (with rate to experimental→ data, described below, the value t−1) will lead to dissociation of the entire motor of b 0.065. As in the forward step, binding is d1 ≈ from the actin filament, terminating the run. followed by Pi release and a power stroke (5 1) −1 → In the kinetic network described above, mechanical occurring at a fast rate tps . forces enter in two ways: the first is through the inter- 3. Leading stomp (10 6 10): This pathway is ini- → → nal strain that leads to the gating mechanism, and the tiated when the leading head detaches with ADP second is through the external load on the junction that still bound (10 6). The reason for this alternative → affects the diffusive search and hence the first passage head detachment mechanism is an asymmetry that ± times tfp. We have already taken into account the asym- arises from the strain in the myosin V dimer in the metry due to internal strain by using experimentally es- waiting state. The backward tension on the leading timated values for td1 and td2, but the force dependence lever arm in the telemark stance (estimated to be ± of tfp has not been specified. This is precisely what the on around 2.7 pN (Hinczewski et al., 2013)) inhibits coarse-grained polymer description allows us to do. We ADP release by a factor of 50-70 (Kodera et al., take advantage of a separation of time scales: when ei- 2010; Rosenfeld and Lee Sweeney, 2004). Rather ther head detaches, the equilibration time tr over which than releasing ADP and binding ATP to detach the polymer legs relax to an approximately equilibrium from actin, the dominant pathway for a head un- distribution of conformations is fast compared to the first der backward load is direct detachment retaining ± passage times, tr tfp. Brownian dynamics simulations the ADP. This has been observed experimentally in and analytical arguments (Hinczewski et al., 2013) show single-headed myosin V, where backwards loads of that tr . 5 µs, while the fastest first passage times are −1 ± around 2 pN lead to an detachment rate td2 = 1.5 at least t (0.1 ms). A summary of all the time −1 fp s (Purcell et al., 2005) independent of ambient scales in the∼ problem O is shown in Fig. 2. Thus to an ex- −1 ATP and ADP concentrations. We thus take td2 cellent approximation the detached head fully explores as the transition rate from 10 6. The difference → some equilibrium distribution of positions, (r), before between the overall detachment rates for the trail- reaching either of the binding sites. In this caseP the first −1 ing and leading heads, with td1 eight-fold larger passage time to reach the forward (+) or backward (-) −1 et al. than td2 , underlies the gating (Purcell , 2005; binding site at position r± is inversely proportional to et al. Veigel , 2005, 2002) mechanism: the trailing (r±), the probability density of finding the detached head is much more likely to detach first. Since de- Phead at that position (Hinczewski et al., 2013): tachment of the leading head occurs with the ADP ± 1 still bound, the head can directly reattach back to tfp , (60) ≈ 4πD a (r±) the forward binding site, with one caveat: the post- h P powerstroke configuration creates an energy barrier Here Dh is the diffusion coefficient of the detached head to reattachment due to geometry (since the lever (which can be estimated as D 5.7 10−7 cm2/s from h ≈ × 27 the crystal structure using HYDROPRO (Ortega et al., backward steps and leading stomps have a mean dura- 2011)). The capture radius a 1 nm is the distance be- tion tLb associated with how long the leading head takes tween the head and binding site≈ for which the interactions to reattach. These two timescales are given by: become strong enough that binding occurs, comparable + + to the Debye screening length under physiological con- t t t = t + fp , t = fp (61) ditions. Eq. (60), which can be derived from standard T b h 1 + bα Lb b + α first passage time analytical approaches like the renewal where α t+ /t− . Note that t is bounded from be- method (Van Kampen, 2007), is extremely useful: it con- ≡ fp fp T b verts the dynamical problem of finding mean first pas- low by th, because hydrolysis is a necessary intermediate sage times of a complex diffusion process into the more step after trailing head detachment, whereas it is not in- tractable problem of calculating the equilibrium distri- volved after leading head detachment. The probabilities bution (r) of the end-point of a two-legged semiflexible that the motor takes a forward step ( f ), trailing stomp P ( ), leading stomp ( ) and backwardP step ( ) are: polymer system. The latter can be found analytically PT s PLs Pb by extending an earlier mean-field theory for semiflexible g t2 polymers (Thirumalai and Ha, 1998), incorporating the = d1 , Pf 1 + g (1 + bα)(t + t )(t + t t ) orientational constraint of the bound leg due to the post- d1 h d1 T b − h powerstroke conformation of the lever arm with respect 1 btd1 (62) T s = bα f , Ls = , to the motor head. Only a handful of structural parame- P P P 1 + g (b + α)(td1 + tLb) ters enter into the theory, determining (r): the contour −1 b = b α Ls, length L = 35 nm and persistence lengthP l 310 nm of P P p ≈ the polymer leg; the angle of the orientational constraint where g td2/td1 = 8 quantifies the strength of gat- ◦ ≡ with respect to the actin filament, θc 60 ; and a pa- ing. These are plotted as a function of load force for the ≈ rameter νc describing the strength of the orientational substall regime in Fig. 15B, along with the termination constraint. The first three are all known from earlier ex- probability t = 1 f T s Ls b. Forward steps perimental estimates (Craig and Linke, 2009; Dunn and predominateP at small− P forces,− P but− P are overtaken− P by trailing Spudich, 2007; Moore et al., 2004). νc is along with b stomps as F approaches the stall value F 1.9 pN, stall ≈ one of the two free parameters in the entire theoretical defined as when f = b. The force dependence of stomp description, and can be estimated based on fitting to ex- probabilities hasP not yetP been measured, and so consti- perimental data. tutes a prediction of the theory, but there is experimen- The full expression for (r) and its derivation can be tal data on the backward-to-forward ratio / (Kad P b f found in Hinczewski et al. (2013). The contours in Fig. 14 et al., 2008). This is shown in comparisonP to theP theo- illustrate (r) for two different values of backwards load retical curve in Fig. 16A, and exhibits excellent agree- P force: A) F = 0 pN; B) F = 2 pN. Superimposed are ment. Other experimentally observable quantities are sample trajectories of the detached endpoint, correspond- compared in the remaining panels of Fig. 16: B) the ing to a forward and backward step respectively, along mean run length zrun along the actin before termination, with snapshots of the polymer conformation at a time from (Baker et al., 2004; Clemen et al., 2005; Pierobon shortly after detachment. The actin filament lies along et al., 2009; Sakamoto et al., 2000); C) the mean velocity the z-axis, with z = 0 corresponding to the location of at- vrun = zrun/trun, where trun is the mean run duration, tached head, and the forward/backward binding sites for from (Clemen et al., 2005; Gebhardt et al., 2006; Kad the detached head at z± = ∆. At zero force the peak of et al., 2008; Mehta et al., 1999; Uemura et al., 2004b). ± the distribution is at z > 0 due to the post-powerstroke The theoretical expressions for these are: orientational constraint on the bound leg. Thus the end-   point probabilities are biased toward the forward binding ∆ 1 α zrun = vruntrun, vrun , site and (r ) (r−). When F = 2 pN the situation ≈ t 1 + bα − g(b + α) P +  P d1 is reversed: the load force pulls the junction backwards, 2 gtd1 counteracting the post-powerstroke constraint, and the trun . ≈ tLb + gtTb distribution is shifted such that (r−) (r+). The (63) dependence of (r) on load forceP translates P into corre- P ± sponding changes in tfp through Eq. (60). The theory curves in Fig. 16 are simultaneous best-fits ± Once tfp as a function of F is known, a variety of phys- with only two free parameters, νc and b, and overall show ical quantities can be calculated directly from the kinetic that the theory quantitatively captures the main features network model (Hinczewski et al., 2013). For example of the motor dynamics (within experimental uncertain- both forward steps and trailing stomps have the same ties evident in the scatter of data points collected under mean duration: the average time tT b from the detach- different buffer and ATP condtions). Equally important, ment of the trailing head to its subsequent reattachment the theory provides insights into how motor properties at either the forward or backward site. Similarly both depend on both kinetic parameters (the gating ratio g, 28

(see (Hathcock et al., 2019) for a fuller discussion). A

VIII. SIMULATIONS USING COARSE-GRAINED MODELS

Several models that can be simulated readily (Alhad- eff and Warshel, 2017; Goldtzvik et al., 2018; Hyeon and Onuchic, 2007a,b; Mugnai and Thirumalai, 2017; Mukherjee et al., 2017; Mukherjee and Warshel, 2013; Nam and Epureanu, 2016; Tehver and Thirumalai, 2010; B Zhang et al., 2017; Zhang and Thirumalai, 2012) have been introduced in order address a variety of issues re- lated to stepping kinetics. These include but are not restricted to the mechanism of stepping of conventional kinesis on microtubule, gating mechanism, and allosteric transitions by which the motor heads communicate with each other and the cytoskeletal filaments. The simula- tions have to be based on coarse-grained models (Hyeon and Thirumalai, 2011) because of the long time scales and interplay of multiple length scales involved in the motor motility. Rather than survey the findings in all these studies, which vary greatly in both details and foci, we describe a particularly illuminating minimal CG mechanochemi- cal model (Craig and Linke, 2009) for myosin V, which beautifully illustrates strain-mediated gating mechanism (required for maintaining processivity) as well as char- acteristics such as speed and stall force. The CG model C (see Fig. 17) , which takes the architecture of myosin V in account, was constructed using the following assump- tions. First, the level arm is treated as a semiflexible polymer by representing the six IQ motifs by three inter- FIG. 16 Best-fit theoretical results (solid curves) from the acting moieties. In this sense the model is similar to the coarse-grained polymer theory for myosin V (Hinczewski et al. et al., 2013) compared to experimental results (symbols) for subsequent analytical polymer model (Hinczewski , the following dynamical quantities: A. the ratio of backward 2013), described above. Second, the junction between to forward steps, Pb/Pf ; B. the mean run length zrun on the head and the adjacent IQ motifs (points 2 and 8 in actin before detachment; C. mean velocity vrun. The sources Fig. 17) was treated as a semiflexible joint with equilib- of the experimental data are listed in the legends (Baker et rium angles that depend on the nucleotide state of the al., 2004; Clemen et al., 2005; Gebhardt et al., 2006; Kad et motor. Such an assumption is justified by comparison al., 2008; Mehta et al., 1999; Pierobon et al., 2009; Sakamoto to EM images. The forward rotation of the lever arm, et al., 2000; Uemura et al., 2004) with buffer conditions in parentheses (first value is ATP concentration, second value is which changes the angle from ΘA to ΘB, is taken to be KCl concentration). Where the KCl value is not listed, the dependent on phosphate release, a crucial step in the re- concentration is 25 mM. All experiments / theory are for satu- action cycle of myosin V. Third, the joint between the two rating ATP conditions except (Gebhardt et al., 2006) in panel lever arms is assumed to fully flexible, which would im- −1 C, where the theory has a modified td1 detachment rate to ply that the tethered head diffuses freely about this joint account for low ATP. Adapted from Hinczewski et al. (2013). (see also (Hinczewski et al., 2013)). It should be noted that recent experiments suggest that this may not be the case. It has been pointed out that the angle between the the reduction in binding rates described by b due to un- lever arms is constrained (Andrecka et al., 2015), which favorable lever arm conformations), and structural pa- has to be taken into account in describing the diffusive rameters (L, lp, θc, and νc that control the distribution search (Hathcock et al., 2019). Fourth, the filamentous during the diffusive search). This allows clear connec- actin is treated as a passive one dimensional track with tions to mutation experiments that perturb the latter, binding sites that are space ∆ = 36 nm apart, as was for example by extending or shortening lever arm length also assumed in the analytical theory (Hinczewski et al., to change L (Oke et al., 2010; Sakamoto et al., 2005) 2013). More importantly, if the tethered head, which has 29

= Q✏2 AAAB+HicbVDLSgMxFL2pr1pfo67ETbAIrspMFXQjFN24bME+wKklk2ba0MyDJCPUofgrbkTcKPgV/oJ/Y6adTVsPBA7nnHDvPV4suNK2/YsKK6tr6xvFzdLW9s7unrV/0FJRIilr0khEsuMRxQQPWVNzLVgnlowEnmBtb3Sb+e0nJhWPwns9jlk3IIOQ+5wSbaSedeQGRA8pEWljct1wWay4MHrnsdqzynbFngIvEycnZchR71k/bj+iScBCTQVR6sGxY91NidScCjYpuYliMaEjMmDpdPEJPjVSH/uRNC/UeKrO5Uig1DjwTDJbUy16mfif95Bo/6qb8jBONAvpbJCfCKwjnLWA+1wyqsXYEEIlNxtiOiSSUG26KpnTncVDl0mrWnHOK9XGRbl2k5dQhGM4gTNw4BJqcAd1aAKFF3iDT/hCz+gVvaOPWbSA8j+HMAf0/QfWipL+ X a Q b ˙ WAAAB5HicbVBdSwJBFL1rX2ZfVo+9DEnQk+yaUI9SLz0apCuoyOw4q4Ozs8vM3UAW/0EvEb0U9Hv6C/2bRt0XtQMDh3POZe65QSKFQdf9dQpb2zu7e8X90sHh0fFJ+fSsbeJUM95isYx1J6CGS6F4CwVK3kk0p1EguR9MHua+/8K1EbF6xmnC+xEdKREKRtFKfm8YY+bPBuWKW3UXIJvEy0kFcjQH5R87ydKIK2SSGtP13AT7GdUomOSzUi81PKFsQkc8Wyw5I1dWGpIw1vYpJAt1JUcjY6ZRYJMRxbFZ9+bif143xfCunwmVpMgVW34UppJgTOaNyVBozlBOLaFMC7shYWOqKUN7l5Kt7q0X3STtWtW7qdae6pXGfX6EIlzAJVyDB7fQgEdoQgsYTOANPuHLCZ1X5935WEYLTj5zDitwvv8AaryLYQ== T2 · 1 W = χ η(ηC − η) 1 T2

(cost) 2 AAAB73icbVDLSsNAFL3xWesr6tLNYBHqpiRV0GXRjcsW7APaUibTSTt0kokzN8US+h1uRNwo+Cf+gn9j2mbT1gMDh3POcO+5XiSFQcf5tTY2t7Z3dnN7+f2Dw6Nj++S0YVSsGa8zJZVuedRwKUJeR4GStyLNaeBJ3vRGDzO/OebaCBU+4STi3YAOQuELRjGVerZdIx3kL5iQIlMGr6Y9u+CUnDnIOnEzUoAM1Z790+krFgc8RCapMW3XibCbUI2CST7Nd2LDI8pGdMCT+b5TcplKfeIrnb4QyVxdytHAmEngpcmA4tCsejPxP68do3/XTUQYxchDthjkx5KgIrPypC80ZygnKaFMi3RDwoZUU4bpifJpdXe16DpplEvudalcuylU7rMj5OAcLqAILtxCBR6hCnVgMIY3+IQv69l6td6tj0V0w8r+nMESrO8/YXiOrg== Q 2kBT

AAAB43icbVDLSgNBEOz1GeMr6tHLYBA8hd0o6DGYi8cI5gFJCLOT3mTM7IOZXiGEfIEXES8K/o+/4N84SfaSxIKBoqqG7mo/UdKQ6/46G5tb2zu7ub38/sHh0XHh5LRh4lQLrItYxbrlc4NKRlgnSQpbiUYe+gqb/qg685svqI2MoycaJ9gN+SCSgRScrNToIPFetVcouiV3DrZOvIwUIUOtV/jp9GORhhiRUNyYtucm1J1wTVIonOY7qcGEixEf4GS+45RdWqnPgljbFxGbq0s5HhozDn2bDDkNzao3E//z2ikFd92JjJKUMBKLQUGqGMVsVpj1pUZBamwJF1raDZkYcs0F2bPkbXVvteg6aZRL3nWp/HhTrNxnR8jBOVzAFXhwCxV4gBrUQcAzvMEnfDnovDrvzsciuuFkf85gCc73HzCBip0=

AAAB3nicbVDLSgNBEOz1GeMr6tHLYBA8hd0o6DHoxWMC5gFJiLOT3mTI7IOZXiGEXL2IeFHwk/wF/8ZJspckFgwUVTV0V/uJkoZc99fZ2Nza3tnN7eX3Dw6Pjgsnpw0Tp1pgXcQq1i2fG1QywjpJUthKNPLQV9j0Rw8zv/mC2sg4eqJxgt2QDyIZSMHJSjW3Vyi6JXcOtk68jBQhQ7VX+On0Y5GGGJFQ3Ji25ybUnXBNUiic5jupwYSLER/gZL7elF1aqc+CWNsXEZurSzkeGjMOfZsMOQ3NqjcT//PaKQV33YmMkpQwEotBQaoYxWzWlfWlRkFqbAkXWtoNmRhyzQXZi+RtdW+16DpplEvedalcuylW7rMj5OAcLuAKPLiFCjxCFeogAOENPuHLeXZenXfnYxHdcLI/Z7AE5/sPckKIYw==

AAAB4XicbVDLSgNBEOz1GeMr6tHLYBA8hd0o6DHoxWME84AkhNlJbzJk9sFMrxBCPsCLiBcFf8hf8G+cTfaSxIKBoqqG7mo/UdKQ6/46G5tb2zu7hb3i/sHh0XHp5LRp4lQLbIhYxbrtc4NKRtggSQrbiUYe+gpb/vgh81svqI2Mo2eaJNgL+TCSgRScMqmLxPulsltx52DrxMtJGXLU+6Wf7iAWaYgRCcWN6XhuQr0p1ySFwlmxmxpMuBjzIU7nG87YpZUGLIi1fRGxubqU46Exk9C3yZDTyKx6mfif10kpuOtNZZSkhJFYDApSxShmWV02kBoFqYklXGhpN2RixDUXZI9StNW91aLrpFmteNeV6tNNuXafH6EA53ABV+DBLdTgEerQAAEjeINP+HKE8+q8Ox+L6IaT/zmDJTjff/Piiec= C ✏X (error) 0

FIG. 18 a. Cost-error tradeoff relation and its physical 2 bound, Q = QX > 2kB T . Accessible region for Q is in cyan. b. Power-efficiency tradeoff. Accessible region for power (W˙ ) is demarcated in cyan.

undergone hydrolysis, diffuses close to a binding site it interacts with the binding site with an attractive electro- static interaction in order to complete a step.

A potential energy function based on the mechanical model, which can be simulated using Brownian dynam- ics, is coupled to the catalytic cycle in one of the motor heads. In order to produce realistic dynamics various rates in the cycle were taken from experiments (see Ta- ble 1 in (Craig and Linke, 2009)). The simulations were successful in reproducing the run length distribution and the value of the stall force (fS 2 3 pN). One of the advantages of the CG simulations≈ is− that the stiffness of the head-neck and neck-neck joints, encoded by the terms VHN and VNN could be changed to assess the effect on the motor properties. For example, they discovered that the value of fS depends on the stiffness of the lever arm. It has to be stiff but not overly so in order to reproduce the experimental data, which was later confirmed theo- retically (Hinczewski et al., 2013).

FIG. 17 Illustrating a general strategy to construct and sim- The strategies used in the models described in this sec- ulate a coarse-grained model for myosin V (Craig and Linke, tion and the previous one are the following. First, the do- 2009). The strategy involves making a coarse-grained model based on the structure (see (A) and (B)), which is coupled mains that execute mechanical movements are modeled to the enzyme chemistry given in (C). (A) The lever arm is using available structural data. Second, the mechanical represented using three rigid segments that are connected to model is coupled to the catalytic cycle, which allows one each other. The two lever arms meet at point 5, which rotates to predict the dependence of measurable quantities as freely during the stepping process. (B) The angle between the a function of control parameters such as ATP concen- lever arm and the head (points 2 and 8) is assumed to change tration and external load. The level of coarse-graining in from ΘA to ΘB upon phosphate release. (C) The reaction the first step is largely guided by intuition. In Hinczewski cycle with various rates indicated in the figure is coupled to et al. the mechanical model. Brownian dynamics simulations were (2013) the use of polymer representation afforded used to calculate the observable quantities. The figure was analytic solution whereas by discretizing the level arm reproduced from Craig and Linke (2009). using discrete connected links Craig and Linke (Craig and Linke, 2009) had to resort to numerical simulations. It is this general strategy that is likely to be successful in tackling the nuances of dynein stepping and perhaps motor functions in vivo. 30

IX. COST-PRECISION TRADE-OFF AND EFFICIENCY greater the heat dissipated from the process, the smaller OF MOLECULAR MOTORS is the error. For molecular motors with output observable x(t), Eq.64 can be written in terms of three quantities 1. Cost-precision trade-off and its physical bound of molecular (heat dissipation rate Q˙ , diffusivity D, velocity V ) that motors depend on control parameters such ATP concentration and external load. From the perspective of thermodynamics, biological systems are clearly in non-equilibrium, which means en- δx(t)2 2D = Q(t) h i = Q˙ 2k T. (65) ergy is constantly injected and dissipated as heat. Be- Q × x(t) 2 V 2 ≥ B cause they are subject to incessant thermal and non- h i thermal fluctuations, cellular processes are inherently Notice that depends on a specific type of motor as well stochastic and error-prone. Thus, a plethora of energy- as the conditionsQ of [ATP] and f. consuming machineries have evolved to fix any error Since the Barato and Seifert’s original conjecture, there that may be deleterious to biological functions. In the has been impressive progress in the field. TUR has presence of large fluctuations inherent to cellular pro- been reinterpreted as the inequality relation between cesses, harnessing energy into precise motion and sup- generalized current and total entropy production rate pressing the uncertainty are critical for accuracy in cel- (σtot = dS/dt), which can be written as lular computation. Trade-off relations between the ener- getic cost and information processing have been a recur- Var(j) σtot 2. (66) ring theme for many decades in biology (Alberts et al., j 2 ≥ 2008; Banerjee et al., 2017; Bennett, 1982; Ehrenberg and h i Blomberg, 1980; Hopfield, 1974; Lan et al., 2012; Mehta General and elegant proofs for TUR have not only been and Schwab, 2012). Recently, a concise and fundamen- given for the case of Markov jump processes on kinetic tal relationship relating cost-precision trade-off and its networks by employing the large deviation theory (Gin- physical bound, which is called the thermodynamic un- grich et al., 2016), but also can be deduced from the certainty relation (TUR), was first conjectured by Barato equality relation for the Fano factor of entropy produc- and Seifert (Barato and Seifert, 2015) and extensively tion for over-damped Langevin processes (Pigolotti et al., studied in the statistical physics community. 2017). Barato and Seifert (Barato and Seifert, 2015) formu- More recently, Dechant and Sasa (Dechant and Sasa, lated the TUR, such that a product ( ) between the heat 2018b) have generalized Eq.66 to underdamped processes dissipation (Q(t)) and the square ofQ relative error asso- in the following form ciated with a time-intergrated output observable X(t) of 2 2 2 σtot B j 2. (67) the process, X (t) = δX / X , is independent of mea- ≥ h i surement time. Basedh oni numericalh i results that exten- where B(> 0) is a model dependent parameter, which sively sampled the rate constants kij defining the diverse kinetic networks and the linear response theory. They can be reduced to 2/Var(j) for over-damped Langevin systems or Markov jump processes on networks. In fact, further conjectured that cannot be smaller than 2kBT for any chemical kineticQ network described by Markov right hand side of the inequality is always greater than jump processes, which is succinctly written as, zero, namely, it recovers the second law of thermody- namics σtot 0. Therefore, one interpretation of Eq.67 2 ≥ = Q(t) X (t) 2kBT. (64) is that it provides tighter bound to the entropy produc- Q × ≥ tion in terms of the square of generalized current. Al- The trade-off parameter quantifies the energetic cost though B is not specific but model dependent, Dechant Q for a given error and is bounded below by 2kBT . The et al.(Dechant and Sasa, 2018b) employed the above re- time-integrated output observable X(t) can be selected lation to derive a power-efficiency trade-off for such that it can best represent the dynamic process of operating between two heat sources with temperature interest. For enzyme reactions that catalyze substrate T1 > T2. to product, X(t) could be the product concentration, 2 c(t). For molecular motors moving along one-dimensional ˙ T1 W χ1 2 η(ηC η). (68) track, displacement (travel distance) x(t) is a natural out- ≤ T2 − put observable to represent their dynamic processes. For motors, the relative error associated with where η = W/˙ Q˙1 and ηC = 1 T2/T1 are the thermo- the motor displacement decreases with time as dynamic efficiency and Carnot efficiency.− The inequality p (δx(t))2/ x(t) 1/√t. Thus, if one were to decide gives the upper bound to the power generated using the h i ∝ the displacement of a motor precisely, a longer time trace two heat sources. But, when η approaches to ηC , W˙ ap- should be generated, which demands more free energy proaches to 0 as well if the model-dependent parameter injection (ATP hydrolysis) and heat dissipation. The χ1 is finite. 31

For pedagogical purpose, it is worthwhile to consider 2. Transport Efficiency simple examples that can demonstrate the significance of the physical bound of TUR. There are a number of ways to assess the “efficiency” (i) In the context of the foregoing one-state hopping of or machines (Brown and Sivak, 2017). Histor- model to which dynamics of molecular motors can be ically, the efficiency of heat engines has been discussed in mapped, the rate of heat dissipation from the process is terms of the thermodynamic efficiency, the aim of which bounded as, is to maximize the amount of work extracted from two heat sources with different temperatures (Callen, 1985). Q˙ u 2(u w)2 = (u w) log − (69) For non-equilibrium machines driven by chemical forces kBT − w ≥ (u + w) that are constantly regulated without shortage in the live cell, the power production could be a more pertinent where the inequality was discussed in Ref. (Shiraishi quantity to maximize. Meanwhile, for transport motors et al. , 2016). Given a single step displacement d0, the in the cell, the TUR parameter can be used to assess velocity and diffusivity of the particle moving along the efficiency of suppressing theQ uncertainty in dynami- the reaction coordinate is V = d0(u w) and D = 2 − cal process by means of energy consumption, and thus is (d0/2)(u + w), respectively. Therefore, inserting the ex- ˙ quite pertinent for evaluating the transport efficiency of a pressions of Q, V , and D to Eq.65 it is easy to see that motor (or motors) (Dechant and Sasa, 2018a). The con- 2kBT . The lower bound of in this model is at- nection between and the transport efficiency is clear. If tainedQ ≥ when u = w, which correspondsQ to the detailed Q ∗ a motor transports cargos at a high speed (V x(t) /t) balance (equilibrium) condition. Provided that u = u [S] 2 ∼ h i ∗ with small fluctuations (D δx(t) /t) (which leads to and v = v [P ], namely, the forward and backward rate punctual delivery to a target∼ site)h buti consuming only a constants vary with the substrate (S) and product (P ) small amount of energy (Q˙ ), such a motor would be con- concentrations, the detailed balance condition is attained ∗ ∗ sidered efficient for cargo transport. A motor efficient in when [S]/[P ] = [S]eq/[P ]eq with u /v = [P ]eq/[S]eq. the cargo transport would be characterized by a small (ii) Hyeon et al. (Hyeon and Hwang, 2017) studied the Q with its minimal bound 2kBT , or as originally suggested over-damped Lagevin motion on tilted washboard poten- by Dechant and Sasa (Dechant and Sasa, 2018a), one tial, which obeys can consider using the definition η = 2k T/ which is T B Q γx˙(t) = f U 0(x) + ξ(t) (70) bounded between 0 and 1. − can be used to assess the “transport efficiency” of Q with U(x + L) = U(x) and ξ(t) = 0, ξ(t)ξ(t0) = biological nanomachines and to study how it changes 0 h i h i 2γkBT δ(t t ), where the driving of quasi-particle on the with varying conditions of f and [ATP]. To evaluate potential is− controlled by the non-conservative force f. , measurement should be first carried out for Q˙ , D, Q They showed that TUR parameter (f) for this problem and V (see Eq.65). While V and D are straightfor- Q is a non-monotonic function of f, attaining its physical ward to calculate (V = limt→∞ dx(t)/dt and D = 0 2 bound 2kBT at both f U (x) (near equilibrium) and limt→∞(1/2)d(δx(t)) /dt), experimental measurement of f U 0(x) (far from equilibrium). | | For f U 0(x) , the ˙ may be nontrivial. Although there are some reports Q quasi-particle | | undergoes diffusion on a rough surface. | | On on direct measurements of heat dissipation rate at single the other hand, for f U 0(x) the particle slides along cell level (Rodenfels et al., 2019; Song et al., 2019), di- a smooth gradient, without | feeling| the effect of confin- rect measurement of heat dissipation at single molecule ing potential, dissipating energy with a rate Q˙ = γV 2 level is not yet known. Nevertheless, Q˙ be estimated by against friction. Since the diffusivity of particle in this considering a physically suitable minimal kinetic network case follows the Stokes-Einstein D = kBT/γ, together model. For a given cyclic kinetic network defined by mul- with the velocity V one can obtain = 2k T . (f) tiple chemical states (i = 1, 2,...N) and transition rates Q B Q reaches its maximum value near the critical point where ( kij ) connecting them, there are straightforward meth- { } the potential barrier confining the particle is about to ods (Hwang and Hyeon, 2017; Koza, 1999; Lebowitz and vanish. Spohn, 1999) to associate the measured V and D with (iii) As far as the specific models are concerned as in kij . As long as all the rate constants kij defining the { } { } (i) and (ii), the minimal uncertainty condition ( min = 2 kinetic network are known, it is then straightforward to Q ˙ kBT ) is attained not only under the detailed balance calculate Q value as discussed in details in the early part condition but when there is no confining potential. It of this review. has been suggested that min is attained when the heat For conventional kinesin, whose chemomechanical dissipated from the processQ is normally distributed as properties were extensively studied by several groups, 2 2 P (Q) e−(Q−hQi) /(2hδQ i) and that TUR measures the motility data for varying [ATP] and load conditions are deviation∼ of heat distribution from Gaussianity (Hyeon available in the literature. The double-cycle network and Hwang, 2017). model, depicted in Fig.8b, with the form of rate con- stant for each edge can be used to analyze those data, 32

FIG. 20 Two case studies for TUR. a. One-state hopping model with forward and backward rate constants u and w. b. Brownian motion in tilted wash-board potential. Q(f) was evaluated as a function f using a specific periodic function, U(x) = U0 sin 2πx/L with L = 6 nm for U0 = 5 and 10 kB T .

replete with obstacles such as cytoskeletal filaments and road blocks). The second local minimum is the very point at which the transport efficiency defined in terms of is optimized. Q (iii) The high region ( > 100 kBT ) at f 3 7 pN is due to theQ stall condition.Q As explained in≈ detail− in rationalizing the double-cycle network model, energy should be still consumed and heat should be dissipated ˙ FIG. 19 a. Analysis of motility data to determine the pa- (Q > 0) at stall condition (V 0). This particular condition renders divergent at≈ stall conditions. rameters for rate constants ({kij } in Eq.47) for double cycle Q network model. b. Diagram of Q as a function of [ATP] and Finally, the structure of (f, [ATP]) is sensitive to Q f. the design of motor structure as well as the motor type. First, the diagram of (f, [ATP]) for a kinesin construct, Kin6AA, whose necklinkerQ is engineered longer than that which allows one to determine the dependence of kij on of wild-type kinesin-1 via insertion of six amino-acids { } [ATP] and f and finally to build a diagram of (f, [ATP]) (AEQKLT) (Clancy et al., 2011) exhibits great devia- Q as shown in Fig.19. The TUR diagram, ([AT P ], f) tion from that of the WT (Fig.21a). In all, the values of Q (Fig.19), exhibit several features that are worthwhile to were increased, the stall forces were reduced, and the explore: localQ minimum observed in the WT is missing. KIF17 (i) In terms of the load direction, assisting (f < 0) and KIF3AB (Milic et al., 2017) show qualitatively sim- and hindering (f > 0), ([AT P ], f) is asymmetric. This ilar structure of (f, [ATP]); however there are different Q result differs from that of the one-state hopping example in terms of quantitativeQ details from that of WT. Next, for TUR calculated in Fig.20a. The fundamental differ- myosin V, dynein, and F1-ATPase were analyzed to cal- ence arises from the asymmetric effect of the force on the culate (f, [ATP]). Bierbaum and Lipowsky (Bierbaum kinetic rate, particularly on k and k with θ = 1/2. and Lipowsky,Q 2011) employed a tri-cyclic network model 25 52 6 (ii) is locally minimized first at a condition of small to describe the chemomechanics of myosin V in which Q load (f & 0) and low [ATP] (indicated by cyan ellipse), cycles for forward steps, energy-consuming futile steps, and second at f 3 pN and [ATP] 300 µM (indicated and force-induced mechanical slippage steps were consid- by the yellow arrow).≈ The first minimum≈ is closer to 2 ered. The resulting (f, [ATP]) (Fig.21d) shows ATP- Q kBT bound; yet, this minimum was attained near the insensitive stall condition where is divergent with no detailed balance condition where [ATP] concentration is local minimum as in WT kinesin-1.Q For dynein (Fig.21e), small and balanced with [ADP] and [Pi]. It does not the uni-cyclic chemomechanical network model adopted have much of biological relevance given that molecular for construction of (f, [ATP]) is in principle not satis- motor works out of equilibrium. In fact, the second local factory, since it wouldQ be able to account for the phys- minimum is of particular interest given that the condition ically correct behavior of dynein dynamics at stall and is not so far from the cellular condtion [ATP] 1 mM superstall conditions. In particular, (f, [ATP]) diagram ≈ Q and f 1 pN, indicated by yellow arrow (Here, note that shows minimization to 2kBT at the stall. Nevertheless, f 1≈ pN is a rough estimate of cellular environment a suboptimal local minimum is found at f 3 pN and ≈ ≈ 33

/k T QAAAB73icbVDLSgNBEOz1GeNr1aOXwSB4irsi6DHEi8cE8oIkLLOT2WTI7MOZ3kBY8h1eRLwo+Cf+gn/jbJJLEgsGiqoauqv9RAqNjvNrbW3v7O7tFw6Kh0fHJ6f22XlLx6livMliGauOTzWXIuJNFCh5J1Gchr7kbX/8lPvtCVdaxFEDpwnvh3QYiUAwikbybLsXUhwxKrP67HbsVRueXXLKzhxkk7hLUoIlap790xvELA15hExSrbuuk2A/owoFk3xW7KWaJ5SN6ZBn831n5NpIAxLEyrwIyVxdydFQ62nom2S+nV73cvE/r5ti8NjPRJSkyCO2GBSkkmBM8vJkIBRnKKeGUKaE2ZCwEVWUoTlR0VR314tuktZd2TW8fl+qVJdHKMAlXMENuPAAFXiGGjSBwQTe4BO+rBfr1Xq3PhbRLWv55wJWYH3/AeskjwQ= B 2

) 7.2 T )

B 7.7 T k (

B 9.1 k AAAB+XicbVDLSsNAFJ34rPUVdaebwSLUTUmqoMtSNy5b6AuaEibTaTt08mDmRiwh4K+4EXGj4E/4C/6Nkzabth4YOJxzhnvv8SLBFVjWr7GxubW9s1vYK+4fHB4dmyenHRXGkrI2DUUoex5RTPCAtYGDYL1IMuJ7gnW96UPmd5+YVDwMWjCL2MAn44CPOCWgJdc8d3wCE0pE0kyxA+wZEpyWp269de2aJatizYHXiZ2TEsrRcM0fZxjS2GcBUEGU6ttWBIOESOBUsLToxIpFhE7JmCXzzVN8paUhHoVSvwDwXF3KEV+pme/pZLanWvUy8T+vH8PofpDwIIqBBXQxaBQLDCHOasBDLhkFMdOEUMn1hphOiCQUdFlFfbq9eug66VQr9k2l2rwt1ep5CQV0gS5RGdnoDtXQI2qgNqLoBb2hT/RlJMar8W58LKIbRv7nDC3B+P4D0raS5w== Q 9.9 ( AAAB5XicbVDLSgMxFL1TX7W+qi7dBItQN2VGBF2WunFZoS9oS8mkmTY0kxmTO0Ip/QQ3Im4U/B1/wb8xnc6mrQcCh3NOuPdcP5bCoOv+Ormt7Z3dvfx+4eDw6PikeHrWMlGiGW+ySEa641PDpVC8iQIl78Sa09CXvO1PHhZ++4VrIyLVwGnM+yEdKREIRtFKnWdSngxqjetBseRW3BRkk3gZKUGG+qD40xtGLAm5QiapMV3PjbE/oxoFk3xe6CWGx5RN6IjP0i3n5MpKQxJE2j6FJFVXcjQ0Zhr6NhlSHJt1byH+53UTDO77M6HiBLliy0FBIglGZFGZDIXmDOXUEsq0sBsSNqaaMrSHKdjq3nrRTdK6qXiWP92WqrXsCHm4gEsogwd3UIVHqEMTGEh4g0/4ckbOq/PufCyjOSf7cw4rcL7/AIomirc= q 13 physical bound 19 (2kT)

AAAB53icbVDLSgMxFL3js9ZX1aWbYBFclZkq6LLoxmUF+8C2lEyaaUMzyZDcEUvpN7gRcaPg3/gL/o1pO5u2Hggczjnh3nPDRAqLvv/rra1vbG5t53byu3v7B4eFo+O61alhvMa01KYZUsulULyGAiVvJobTOJS8EQ7vpn7jmRsrtHrEUcI7Me0rEQlG0UlPbZ5YIbXqvnQLRb/kz0BWSZCRImSodgs/7Z5macwVMkmtbQV+gp0xNSiY5JN8O7U8oWxI+3w823NCzp3UI5E27ikkM3UhR2NrR3HokjHFgV32puJ/XivF6KYzFipJkSs2HxSlkqAm09KkJwxnKEeOUGaE25CwATWUoTtN3lUPlouuknq5FFyWyg9XxcptdoQcnMIZXEAA11CBe6hCDRgoeINP+PKE9+q9ex/z6JqX/TmBBXjff7SajLo= x

FIG. 21 Q(f, [ATP]) for a. Kin6AA (kinesin-1 mutant); b. FIG. 22 Q values for various transport motors calculated at KIF17; c. KIF3AB; d. myosin V; e. Dynein; f.F1ATPase. [ATP] = 1 mM and f = 1 pN (except for F1-ATPase calcu- The diagrams were calculated by directly analyzing the motil- lated at [ATP] = 1 mM and τ = 0 pN·nm) ity data available in (Clancy et al., 2011) for a, (Milic et al., 2017) for b, c, and using the same double-cycle network model as that of kinesin-1. For other motors in d, e, and f, the ki- X. MOLECULAR CHAPERONES netic network and rate constants were used as in the Ref. (Bierbaum and Lipowsky, 2011) for myosin V, Ref. (Sarlahˇ and Vilfan, 2014) for dynesin, and Ref.(Gerritsma and Gas- Chaperones are another class of cellular molecular ma- pard, 2010) for F1-ATPase. chines that expend free energy change associated with ATP binding and catalysis to facilitate the folding of cer- tain proteins and RNA that cannot fold spontaneously [ATP] 300 µM. Finally, (τ, [ATP]) (torque τ instead under cellular conditions. Because they serve a key func- ≈ Q of load f) for F1ATPase is shown in Fig.21f calculated tion in maintaining protein homeostasis they are deemed based on uni-cyclic kinetic network model (Gerritsma as essential for the survival of the organisms. Among a and Gaspard, 2010). For F1-ATPase, which shows near- large class of molecular chaperones, the function of the reversibility and hence characterized with high efficiency heat-shock GroEL-GroES machinery found in E.Coli, re- (Toyabe et al., 2010), the use of uni-cyclic network model ferred to as cheperonin, is now quantitatively understood would be reasonable although this conclusion should be thanks to experimental and theoretical advances. The in reached with care (see (Sumi and Klumpp, 2019)). vivo function of GroES-GroEL machinery in E. Coli. is Some of the biological motors, kinesin family and to rescue substrate proteins that are otherwise destined dynein, studied here are found to be semi-optimized in for aggregation. Just like molecular motors, GroEL un- terms of under the cellular condition, which alludes dergoes a catalytic cycle involving a series of large scale to the roleQ of evolutionary pressure that has shaped the structural changes in response to ATP binding, hydroly- present forms of molecular motors in the cell. In addi- sis, and release of ADP and phosphate. GroEL/GroES tion, the efficiency quantified in terms of for various machine anneals the population of misfolded proteins molecular machines presented here are rangedQ between 7 driving them to the folded state by repeatedly going and 20 kBT (Fig.22). Given that all these machines func- through rounds of the catalytic cycle, which has been tion out-of-equilibrium, it is of great interest to discover referred to as the Iterative Annealing Mechanism (IAM) that the value of calculated at the working cellular con- (Tehver and Thirumalai, 2008; Todd et al., 1996). Q dition is not significantly far from its physical bound 2 A minimal kinetic network model of chaperonin- kBT . This may arise from the fact that the molecular ma- assisted protein folding, can be constructed, based on chine analyzed here are tightly coupled machine, mean- considerable experimental evidence, by assuming that a ing that ATP hydrolysis is almost always transduced to a protein exists either in an intermediate (I), misfolded machanical step even though energy-wasting futile steps (M), or folded (native) (N) state as illustrated in Fig.23. still remain as a possibility. The unfolded state, a transient state right after the pro- As long as the underlying mechanism, which offers tein is synthesized, collapses to the I state, and it further a clear model for chemomechanical kinetic network, is undergoes a spontaneous folding process via the kinetic known, TUR can be studied for any time trace generated pathways with rates denoted by kIM and kIN . Only a from cyclic process at NESS. Other energy-intensive fraction (Φ) of the entire population folds correctly to the processes, for example, error-correction processes, circa- N state and the remaining fraction (1 Φ) is misfolded dian cycle, and chaperonin action, would exhibit high . to the M state. The process of the initial− population Q of molecules with Φ (1 Φ) reaching the folded (mis- − 34

folded) state is termed the Kinetic Partitioning Mech- anism (KPM). The KPM, which was first theoretically proposed (Guo and Thirumalai, 1995), has been used to explain the folding of of proteins (Kiefhaber, 1995) as well as RNA (Pan et al., 1997; Thirumalai and Hyeon, 2005). Typically, substrate proteins (SPs or ribozymes) that require assistance from chaperonin action are char- acterized by extremely small values of Φ ( 1), which implies that the majority of population without chap- erones are trapped in the misfolded states, which could potentially aggregate unless rescued by the chaperones. The function of the GroEL machine is quantitatively explained by the Iterative Annealing Mechanism (IAM) according to which the chaperone recognizes and acts on misfolded substrate preferentially. Typically, the SPs have exposed hydrophobic residues, which make them prone to aggregation unless recognized by the GroEL- GoES machine. When the SPs bind to GroEL they become disordered as a result of domain movements in GroEL, which imparts a mechanical force ( 10pN) that is sufficiently large to unfold (at least partially)≈ the SPs (Thirumalai and Lorimer, 2001). As the catalytic cy- cle proceeds the SPs are encapsulated for a brief period (roughly about two seconds) during which a small frac- tion folds rapidly during the time they are in the cavity. It is worth remarking that if they fold (the probability be- ing Φ), they do so while being encapsulated in the cavity of GroEL. When the catalytic cycle is complete, the SP is ejected from the cavity regardless of whether it is folded or not. If the SP is misfolded then it once again recog- nized by GroEL and the cycle is iterated until sufficient yield of the folded state is obtained. A key requirement of FIG. 23 Chaperone-assisted folding of substrate. a. GroEL-SP interaction is that hydrophobic residues of the Schematic of the iterative annealing mechanism (IAM) illus- SPs must be exposed, which does not typically occur in trated for the hemicycle of GroEL. The T R transition starts when ATP and the substrate protein (SP) bind. GroES folded states. It this ability of chaperone not to recognize binding and ATP hydrolysis engenders the R0 → R00 transi- native proteins that enables the GroEL-GroES machin- tion with a fraction of the SP partitioning to the folded struc- ery to drive the misfolded states to the native state over ture. Subsequently ADP and Pi and the SP (folded or not) repeated iterations of the catalytic cycle. More specifi- 00 are released and the R → T transition completes the cycle. cally, when the annealing process, corresponding to one In the presence of the SP the machine turns over in about catalytic cycle, is iterated n times, the total amount of a second and ADP release is accelerated by about a hundred fold. The rapid turnover comports well with the predictions of population that reaches the native state (or native yield) IAM (Todd et al., 1996). b. A minimal kinetic network model grows as, for chaperone-assisted folding of a substrate molecule. Tran- n sitions between three manifolds of collapsed intermediate (I), Nn = 1 (1 Φ) (71) misfolded states (M), and native state (N) are represented − − in terms of rate constants. Folding, misfolding, and chaper- with n 1. As n , Nn 1. one assisted unfolding pathways are depicted in blue, red, and ≥ → ∞ → While it is known that GroEL only recognizes mis- magenta, respectively. c. Schematic of folding landscape of substrate molecules. Upon spontaneous folding, the ensemble folded SPs, the IAM concept has to be generalized to of intermediate states collapsed from unfolded ensemble reach RNA chaperones, which act on the both N and M states, the native and misfolded basins of attractions with the pro- but more favorably on M (Bhaskaran and Russell, 2007; portion of Φ and 1−Φ, respectively. d. Schematic of the gen- Chakrabarti et al., 2017). If the proportion of N iden- eralized iterative annealing mechanism of chaperone-assisted tified by chaperones in comparison with the M state is substrate folding, from which the recursion relation for the defined as κ (0 κ 1), the total yield of native state native yield from n-th annealing process was derived. N and i after n rounds of≤ folding≤ (annealing) in the presence of Mi denote the proportion of native and misfolded states from the i-th annealing process. The blue and red arrows repre- chaperone can be calculated using the following mathe- sent the pathways leading to the native and misfolded states, matical formulation: respectively. 35

ss ss (i) Let Nn and Mn be the yields of native and misfolded (J = kijPi kjiPj ) can be expressed solely using the state, respectively, after the n-th round of the annealing rate constants,− as already discussed in depth in the fore- process. Note that the total amount of substrate proteins going section describing the general aspects of the vari- is conserved at all times, which implies that Nn +Mn = 1 ous cycles. Because the function of the chaperones is to for all n. In the first round of annealing, a fraction N1(= promote the formation of the folded state our primary in- Φ) folds to the N state and M1(= 1 Φ) partitions to terest is the steady state yield of the native state, which the M state; − is given by, (ii) In the n-th round of annealing, chaperones rec- ss kMI ([C], [T ])kIN + kMN (kIM + kIN ) ognize Nn−1 and Mn−1 differentially by the factor κ. P = , (75) N Σ([C], [T ]) Whereas (1 κ)N − is left unrecognized by chaper- − n 1 ones, κNn−1 is unfolded and Φ of them refold to yield where κΦNn−1 native states and κ(1 Φ)Nn−1 misfolded states. − Σ([C], [T ]) = On the other hand, the entire population of Mn−1 is un- folded and Φ of them refold to yield ΦMn−1 native states kMI ([C], [T ])kNI ([C], [T ]) and (1 Φ)Mn−1 misfolded states. Therefore, after the + kNI ([C], [T ])(kMN + kIM ) + kMI ([C], [T ])(kIN + kNM ) (n 1)-th− round of chaperone action, the native yield N n + (k + k )(k + k ). is determined− as N = 1 M = 1 (1 Φ)M κ(1 IM IN NM MN n n n−1 (76) Φ)N ; − − − − − n−1 The partition factor Φ can be expressed in terms of rate (iii) From the resulting recursion relation of N = (1 n constants as Φ = k /(k + k ). Because of chaper- κ)(1 Φ)N + Φ with N = Φ, we obtain the following− IN IN IM n−1 1 one action, k and k can be significant (in particular expression,− IN MI k k ), whereas k and k are negligible. Under IN  NI NM MN 1 (1 κ)n(1 Φ)n this condition, the native yield simplifies to: Nn = Φ − − − , (72) κ + (1 κ)Φ ss kMIkIN − PN native yield after n iterations. After a sufficient number ' kMIkNI + kNIkIM + kMIkIN k Φ of iterations (n ), the system reaches steady state, IN = = N (77) → ∞ k ∞ N∞ Φ/(κ + (1 κ)Φ). In the case of GroEL only ' NI κ + (1 κ)Φ → − kIM + kIN − the misfolded state is recognized by chaperones (κ = 0). kMI Therefore, the GroEL-GroES machinery drives the en- In the absence of either of chaperone or ATP which tire population of substrate proteins to the native state redistributes the population of proteins into NESS, [C], N∞ 1 (Eq.71). → [T ]-dependent rate constants vanish (kij([C], [T ]) = 0). As described for molecular motors, the thermodynamic In this case, the steady state population of N state be- aspects of chaperonin-assisted folding can be succinctly comes captured by mapping the folding process of substrate ss 1 molecule onto a kinetic network model. A uni-cyclic re- PN ([C] = 0 or [T ] = 0) = versible kinetic network model consisting of the above- 1 + kNM /kMN 1 mentioned three states (I, M, and N) suffices to capture = the non-equilibrium nature of chaperone-assisted protein 1 + e−∆GNM /kB T eq folding. The relevant master Equation describing the = PN (78) network is, Replacing the two [C], [T ]-dependent rate constants ∂tP(t) = P(t) (73) (kIN and kIM ) to zero is tantamount to blocking the W chaperone action and placing the M and N states in T where P(t) = (PI (t),PN (t),PM (t)) and, isolation. As long as they are isolated for long enough −1 −1   time greater than kNM and kMN , the system would (kIN + kIM ) kNI kMI eq − finally reach the equilibrium native yield (PN ), as dic- =  kIN (kNI + kNM ) kMN  tated in Eq.78 assuming that aggregation reaction can W k − k (k + k IM NM − MI MN be neglected. However, equilibrating M and N states via the transition paths of M N is impractical in the The probability Pi(t) of each state i evolves with time −1 light of the time scale of cellular processes because kNM as, −1 and kMN would far exceed the biologically meaningful ss −λ1t −λ2t P(t) = P + c1~u1e + c2~u2e (74) time scale. These arguments suggest that chaperones drive the substrates out of equilibrium. In the process, with λ2 > λ1 > 0, and it reaches the steady state they optimize the yield of the folded SPs or per ss value Pi at t . The steady state population unit time, which we discuss further below. ss → ∞ Pi and steady state current along the reaction cycle 36

licases destabilize the base pairs of the dsDNA, perhaps A B by exerting a force, thus separating the two strands. If the helicase is passive then it binds to the ssDNA when- E ever thermal fluctuations transiently open the base pairs. 250 As the strands of the dsDNA are separated the helicase 4000 )

bp 3000 translocates along the ssDNA. This strand separation ( C ) 200 s / 2000 and translocation are intimately related. Several ensem- bp

( 150 1000 Processivity ble experiments have provided glimpses into the stepping 0 100 0 2 4 6 8 10 Force (pN) mechanism (Ali and Lohman, 1997; Jeong et al., 2004; Velocity 50 Levin and Patel, 2002; Lucius et al., 2003; Wong and D 0 0 2 4 6 8 10 12 Lohman, 1992) of helicases. In addition, these experi- Force (pN) ments have also been insightful in deciphering how they interact with single strand–double strand (ss-ds) junc- tions, and how often they dissociate from their track. The most detailed picture of the functions of a number of helicases have come from single molecule laser opti- cal tweezers (LOT) and (MT) exper- FIG. 24 Helicase model. Panels (A-D): Schematic of a model iments. These experiments provided the kinetics of step- showing the unwinding of nucleic acids by helicases. (A) ping and nucleic acid unwinding (Dessinges et al., 2004; + − Translocation process of a helicase, with k , k , and γ being Lionnet et al., 2006; Patel and Picha, 2000; Perkins et al., the rates of forward, backward, detachment, respectively; n is 2004). Such measurements are instrumental in not only the position on the nucleic acid. (B) Opening (α) and closing formulating theories and simulations but also in refining (β) rates of a base pair. (C) Interactions between helicase and nuceic acids modify the stepping kinetics of the helicase and them as additional high precision experiments become and the opening and closing rates of the base pair. The base available. pair location is m and j = m − n. (D) A model for the inter- In a remarkable and influential paper, Betterton and action with U0 being the strength. For a passive helicase U0 is Julicher (BJ) (Betterton and J¨ulicher, 2003, 2005a,b) zero. Panel (E): Simultaneous fits of the force-dependence of presented a theory, which describes quantitatively the the velocity and processivity (given in the inset) of T7 helicase as a function of an external load. The blue and red lines are coupling between translocation and strand separation. the results using the theory described elsewhere (Chakrabarti The framework used in this theory, which is another il- et al., 2019) and the data points in circles are taken from lustration of the SKM, has been most instrumental in experiments (Johnson et al., 2007). understanding the differences between active and passive helicases. The BJ model assumes that the helicase moves forward (backward) at a rate k+ (k−) when it is very far XI. UNIVERSAL CHARACTERISTICS OF HELICASES from the ss-ds junction. This aspect of the motor move- ment is similar to Fig. 5 with γ=0. When the motor Helicases, which are molecular motors found in all or- is very far from the ss-ds junction the helicase merely ganisms, unwind double stranded (ds) DNA and dsRNA translocates along the ss nucleic acids. Similarly, in iso- when they encounter a junction between single strand lation a nucleic acid base opens at a rate α and closes at (ss)-ds nucleic acids (Delagoutte and Von Hippel, 2002, a rate β. Depending on whether it is a AT or a GC base 2003; Lohman, 1992; Lohman and Bjornson, 1996). Sep- pair the rates are different but the inequality β α holds aration of dsDNA strands is required for DNA replication in both cases. Since isolated base pair opening rates are α −∆Gbp/kB T as well as DNA repair. Malfunction of helicases causes due to thermal fluctuations they satisfy β = e genomic instability and are also implicated in cancer. where ∆Gbp is stability of the base pair. However, inter- In addition, certain RNA chaperones are also deemed actions with base pairs modify these rates. Let n be the to have helicase activity, which means they are able to position of the helicase on the track and m be the loca- unwind helices in RNA in order to facilitate its folding tion of the ss-ds junction (Fig. 24). Upon approaching (Mohr et al., 2002; Russell et al., 2013). We refer the the ds-ss junction the helicase interacts with NA, which reader to a number of articles that describe a variety of was modeled using a variety of potentials - all based on cellular functions associated with helicases (Bianco and some combination square well like potentials. Passive Kowalczykowski, 2000; Bustamante et al., 2011; Dong helicases, characterized by U0 , opportunistically et al., 1995; Lohman et al., 2008; Marians, 2000; Pang step when the base is open. For→ active ∞ helicases, which et al., 2002; Pyle, 2008; Rocak and Linder, 2004; Ve- forcibly rupture the base pair interactions, U0 is finite but lankar et al., 1999; Venkatesan et al., 1982). Helicases, depends on j = m n. The rates k+ and k− and α and classified into six super families based on their sequences β are modified when− the helicase interacts with the NA (Gorbalenya and Koonin, 1993; Iyer et al., 2004), are de- (Betterton and J¨ulicher, 2005b). To describe the action scribed as active or passive (Lohman, 1992). Active he- of the helicase one has to keep track of its position on the 37 track as well the ss-ds location. Helicase-NA interactions base pairs. (iii) Normally, in analyzing experimental data modify all the relevant rates making them position (j) back NTP-dependent stepping rates are neglected. This dependent. Using the model, with detachment rate set is justified while the helicase translocates along ss NA. to zero, the velocity of the helicase assuming that it steps For instance, in T7 helicase the ratio of the forward to one base pair at a time, is given by, back stepping rate is 270, a value that is not that different from kinesin at∼ zero resistive force. The proba- 1 X V = k+ + α k− β  P , (79) bility of back-stepping is 0.3%. However, when T7 un- 2 j j − j − j j ≈ j winds dsDNA the probability of back-stepping increases to 26%. This estimate is not that dissimilar to obser- where Pj is the probability of observing the helicase and vations in XPD helicases belonging to a different (SF2) junction that are separated separated by j, αj is the rate family, where it was shown that at 1mM ATP concentra- at which the junction opens when the helicase and junc- tion the back-stepping probability is 10%. (iii) Inter- ∼ tion are at separation j. The corresponding rates for base estingly, many helicases do not function in vitro unless an pair closing, forward, and backward stepping rates of the external force is applied. Under in vivo conditions part- + − helicase are βj, kj , and kj , respectively. ner proteins that bind to single strands and impart a force Although experiments (see for example (Manosas at the ss-ds junction to destabilize the base pairs (Pincus et al., 2010)) have been analyzed using the theory et al., 2015) are needed. That this is the case has been sketched above the BJ model has to be generalized in shown for UvrD helicase (Comstock et al., 2015), which order to make precise comparison with experiments. (1) behaves as a processive motor only when 2pN force is To account for the average helicase processivity ( δ m), applied. It was noted previously that even though these which was not addressed in the original formulation,h i the associated proteins may or may not increase the unwind- consequences of detachment of the motor have to be con- ing velocity of a helicase they should universally increase sidered. As pointed out earlier, almost all of the chemi- the processivity of the helicase. cal kinetics models ignore detachment, which of course is The theories for helicase stepping are not complete be- unrealistic because the run-length in motors or number cause they do not resolve many of the challenging prob- of base pairs that are disrupted if finite. (2) A theory lems. First, there is no quantitative explanation for the that allows for arbitrary step size (s) is needed instead very broad velocity distribution (Johnson et al., 2007) of assuming that s is unity. Helicases, such as PcrA and measured during unwinding of the weakly active ring- NS3 interact with and possibly destabilize several base shaped T7 helicase. It is challenging to calculate P (v) pairs that are down stream of the ss-ds junction (Cheng for the recent model (Chakrabarti et al., 2019), which et al., 2007; Velankar et al., 1999). In other words, step is the minimal that accounts for the F -dependent mean size s exceeds unity. (3) Experiments also apply external velocity and δm accurately. Second, there is very lit- force and measure the changes in the processivity and tle understanding of the structural basis of the universal velocity as a function of force, f. A viable theory should increase in δm as a function of F and the underly- produce tractable expressions for both the mean velocity ing dramatich variationsi in sequence and architectures of and δm as a function of of quantities. (4) Finally, how the motor. Perhaps, carefully designed simulations might h i the sequence of the NA affects δm and V needs to be shed light on this issue (Ma et al., 2018; Yu et al., 2006). h i considered in order to draw general conclusions. Third, a recurring theme in many aspects of motors is An analytically solvable model that accounts for the that the dynamics is heterogeneous, exhibiting charac- first three effects stated above has been proposed recently teristics reminiscent of glasses. It was discovered that (Chakrabarti et al., 2019), which was preceded by a less there are substantial molecule-to-molecule variations in general theory (s was set to unity) (Pincus et al., 2015) the unwinding speed of E. Coli. RecBCD helicase even if that investigate sequence effects ob the mean velocity and all the enzymes are prepared under the same condition. δ m. These studies produced a number of unexpected To account for this observation it has been suggested predictionsh i for the helicase velocity and processivity as a that the functional landscape is likely partitioned into a function of external force and DNA sequence. (i) It was number of metastable states and the initial preparation predicted that, regardless of the underlying architecture quenches the enzyme into a specific substate (Kirkpatrick and unwinding kinetics of the helicase or the precise DNA and Thirumalai, 2015) The helicase ergodically explores sequence, processivity exhibits a universal increase with all the conformations within a single metastable state but applied external force. This finding, which has subse- the transitions to other states could only be achieved by quently has been validated experimentally (Bagchi et al., resetting the ATP concentration (Liu et al., 2013). The 2018; Li et al., 2016), was used to suggest that helicases emergence and relevance of glass-like heterogeneous be- may have evolved to maximize processivity rather than havior, under ambient conditions, is not understood the- velocity. (ii) The theory, which quantitatively accounts oretically, and remains a challenge not only in the context for the experiments for force-dependent V and δm for of helicases but also in other biological systems as well T7 replisome (Fig. 6 from BJ 2019), shows thath si= 2 (Altschuler and Wu, 2010; Hyeon et al., 2012; Solomatin 38 et al., 2010). vations are amenable to theoretical treatments exposed here, in which molecular details are ignored. Detailed simulations, if possible, could provide biophysical insights XII. DISCUSSION but linking such studies to functions is a daunting task. These anecdotal examples should remind us that in the The combination of excellent experiments and few the- search of principles of generality in biology one should oretical approaches touched on here have greatly ad- not forget that molecular details matter and could dra- vanced our understanding of how biological machines matically influence functions. work. However, we still do not have a complete under- standing of how these machines work even in in vitro conditions. This should not be a surprise because some believe that the link between allosteric communication in B. Efficiency and optimality hemoglobin (Hb) and oxygen transport is not fully un- derstood despite over fifty years of intensive study. It A naive assessment of efficiency may be made by es- is unclear if at present there is an analogue of Hb, con- timating the theoretical stall force, based on the avail- sidered the hydrogen molecule for allostery, in molecular able free energy due to ATP hydrolysis, and comparing machines. The investigation of these machines, one at a it to the measured stall force (Kolomeisky and Fisher, time in vitro, seems to raise unsolved problems in each 2007). Consider F1-ATPase for example. This is part individual case. We outline a few of the challenging prob- of the F0F1-ATP synthase responsible for synthesizing lems. Clearly, the list is far from being exhaustive. ATP. This rotary motor undergoes precise 120◦ rotations in the absence of an external applied torque (τtor) at low ATP concentrations. By controlling τtor using the electro A. Sometimes Details Matter rotation method and the chemical potential by choosing appropriate ATP, ADP, and Pi concentrations, it is pos- Here, we have only described coarse-grained theoretical sible to measure the probability (ps) of rotation in the methods, which are impervious to the molecular details. synthetic direction (ADP and Pi are consumed to gener- There are several examples (we mention two) in which ate ATP and the probability ph in the reverse hydrolytic spectacular changes occur by a single or few amino acid direction could be measured (see for example (Toyabe et al., 2011)). From the linear dependence of k T ln ps substitutions. (i) About twenty years ago, Endow and B ph Higuchi (Endow and Higuchi, 2000) made a single amino on it was found that the output energy at stall is roughly acid substitution in the neck linker (NL) region of a Ncd, equal to the chemical potential. This implies that F1- a motor that is related to kinesin. The wild type Ncd ATPase operates at near 100% efficiency. walks towards the minus end of the microtubule in con- For myosin motors, which take roughly a d=36 nm trast to conventional kinesin. Upon substituting an as- step, the maximum force that can be exerted is fmax ∆GAT P ≈ parigine (a polar amino acid residue) to lysine (positively d , which is approximately 2.5 pN assuming that ∆G 22k T . The measured stall force (f ) is charged) in the NL, an element that is responsible for the AT P ≈ B stall motor to walk predominantly on a single protofilament roughly in this ball park, which suggests that myosin motors operate efficiently (η = fstall is very high). A of the MT, it was found that Ncd moves in both the plus fmax and minus direction on the MT. (ii) Myosin VI, unlike similar argument for kinesin yields (d=8.1nm) fmax ≈ myosin V, walks towards the filamentous actin minus end 12pN whereas measurements report values close to 8 pN. and is the only known member belonging to the myosin Thus, there is about a 30-35% decrease in η for kinesin super family with this property. In humans there are re- motors. A precise computation of efficiency should be ports of three mutations that cause deafness. A point undertaken by considering the network for a given motor missense mutation (replacement of aspartic acid by tyro- that captures many aspects of motor motility. The ar- sine (D179Y) in the so-called U50 domain of the motor) guments given above hold roughly if the motility can be leads to deafness in mouse (Hertzano et al., 2008). It described in a periodic one dimensional tilted potential has been suggested (Pylypenko et al., 2015), using a va- with two equivalent sites with a transition state, which riety of experimental methods, that D179Y mutant leads is close to the initial site. Treatments (Golubeva et al., to a premature release of the phosphate Pi, a product 2012; Schmiedl and Seifert, 2008; Seifert, 2011a; Wagoner of ATP hydrolysis, from the detached head, thus pre- and Dill, 2019) using more elaborate models indicate that venting it from rapidly binding to F-actin. In the mean- the the motor efficiency would be much less and would while, ATP does bind to the leading head, which de- depend on the details of the network dynamics. taches the motor from actin, thus preventing processive A question that is related to efficiency is optimal per- motion. Besides these examples there are many others, formance. In the context of the biological machines dis- such as the link between mutations in β-cardiac myosin cussed here, performance should be measured by velocity and hypertrophic cardiomyopathy. None of these obser- of movement, processivity, and for molecular chaperones 39 the time-dependent production of folded state. As men- D. Biological complexity tioned above, studies in the last decade have addressed how biological machines might optimize speed by con- We circle back to Fig. 1, which is a schematic illus- sidering models that are used to analyze force-velocity tration of transport of melanosomes (vesicles containing curves in motors. One of the lessons is that speed might the light absorbing pigment melanin found in amphib- be optimized if the motor takes many sub-steps instead ians). The vesicles are either dispersed throughout the of a single step (Wagoner and Dill, 2016). These studies cytosol or aggregate near the cell center. The transporta- have not considered processivity (run length) as a func- tion of melanosomes is clearly complex and is controlled tion of ATP and external force. For helicases, it appears by the interplay of multiple motors involving kinesin-2, a that maximization of velocity is not as relevant as opti- plus end directed MT motor, and dynein, that walks to- mization of processivity. In addition, for GroEL it ap- wards the minus end of the MT. In addition, actin bound pears that the rate of production of the folded state per myosin V is also involved in the transport. It is suspected unit time is maximized even at the consumption of lav- that a low number of motors (about 1-2 kinesin-2 motors ish amount of energy, which would render this machine and roughly 1-3 ) move melanosomes during ag- highly inefficient. Whether questions pertaining to opti- gregation (Levi et al., 2006). In contrast, pigment dis- mal performce, given available free energy, must consider persion is mediated by kinesin-2 as well as assistance by simultaneously many functional requirements remain an myosin V. Because both these motors compete for the open problem. It is possible that optimality could depend same attachment site (p150Glue on dynactin, see Fig. 1) on specific function carried out by a class of machines. it follows (Levi et al., 2006) that myosin V is released during aggregation and vesicle transport is dominated by dynein. The need to change movement of melanosomes, powered by multiple motors, is thus determined by func- C. Specificity versus Promiscuity tion, which in this case is related to their dispersion or aggregation. The E. coli chaperonin has evolved to be a promiscu- Although individual motors predominantly move uni- ous machine in that it facilitates the folding of a variety directionally on cytosketal filamentous tracks there are of misfolded substrate proteins (even those that are not reports that motors could change directions as well. A in the E. coli proteome) that are unrelated by sequence, very impressive in vitro illustration of bidirectional motil- size, or the structure of the folded state. Using directed ity of the complex of dynein with dynactin (a complex evolution methods a mutated GroEL/GroES, referred to that is attached to the cargo and activates dynein (Fig. 1) et al. as GroEL3−1 was constructed (Wang et al., 2002). The was reported sometime ago (Ross , 2006). It was altered GroEL3−1 contained a single mutation in GroES found that the complex moves processively in both di- (tyrosine was replaced by histidine) and two mutations rections on MT with ATP-dependent velocities that are (valine was substituted for alanine and glycine for aspar- not significantly different in either direction (Ross et al., tic acid) in GroEL. It was found that GroEL3−1 had en- 2006). Although the authors provide a qualitative picture hanced ATPase activity compared to the wild type. More of the mechanism of bidirectional transport a theory for importantly, it was found that GroEL3−1, with a highly such unexpected behavior is lacking. It is not even clear, polar environment in the cavity compared to the wild- at least to us, the level of coarse-graining needed to con- type, dramatically increased the folding of green fluores- struct such a theory, which is clearly needed to unveil the cent protein (GFP). However, the enhanced specificity of complexity of vesicle transport. folding GFP with ease came at the expense of substan- tial reduction in the capacity of GroEL3−1 to facilitate the folding of several other proteins. Thus, in this in- XIII. A FINAL REMARK stance nature has solved the tension between specificity and promiscuity by evolving an all purpose E. coli chap- The eventual goals of understanding biology through eronin that can process the folding of a large class of the lenses of physics are to create theoretical tools proteins, albeit not as efficiently. In eukaryotes there that are capable of describing biological functions under has been a great expansion in the number of chaperone crowded and noisy cellular conditions, and in the process classes possibly to enable the larger and more complex discover general physical principles that control life pro- proteome. This example suggests that evolutionary con- cesses. It is likely that as the scale at which living systems straints might have to a part of addressing issues related are examined increases it may be possible to describe cel- to optimality. This might imply that the simple network lular processes using functional modules (Hartwell et al., used to explain the out of equilibrium performance of 1999), which is a coarse grained view of biology. How- GroEL/GroES has to expanded to describe optimality in ever, it would be hard to anticipate the functions of such the functions of chaperone networks in eukaryotes. modules from their components, which are the molecules of life. Furthermore, interactions between modules could 40 lead to new functions not encoded in isolated modules, as Banerjee, K., A. B. Kolomeisky, and O. A. Igoshin (2017), illustrated by many examples described in our perspec- Proceedings of the National Academy of Sciences 114 (20), tive. After all “more is different” (Anderson, 1972). Be- 5183. cause “Nature is an excellent tinkerer, not an engineer” Barato, A. C., and U. Seifert (2015), Physical review letters 114 (15), 158101. (a quote attributed to Francois Jacob), and tinkering in Battle, C., C. P. Broedersz, N. Fakhri, V. F. Geyer, biology involves stochastically altering existing modules J. Howard, C. F. Schmidt, and F. C. MacKintosh (2016), to evolve new functions without time constraints or any Science 352 (6285), 604. ultimate design as a goal (Jacob, 1977), it is likely that Bell, G. I. (1978), Science 200 (4342), 618. concepts in many fields of science would have to be used Belyy, V., N. L. Hendel, A. Chien, and A. Yildiz (2014a), to develop an integrated view of biology. Nat. Comm. 5, 5544. Belyy, V., N. L. Hendel, A. Chien, and A. Yildiz (2014b), Nat. Comm. 5, 5544. Bennett, C. H. (1982), International Journal of Theoretical ACKNOWLEDGMENTS Physics 21 (12), 905. Betterton, M. D., and F. J¨ulicher (2003), Phys. Rev. Lett. We are grateful to: Matthew Caporizzo, Shaon 91 (25), 258103. Chakrabarty, Yale Goldman, Yonathan Goldtzvik, Betterton, M. D., and F. J¨ulicher (2005a), Phys Rev E 71 (1 Sabeeha Hasnain, William O. Hancock, David Hath- Pt 1), 011904. cock, Chris Jarzynski, Anatoly Kolomeisky, George H. Betterton, M. D., and F. J¨ulicher (2005b), J. Phys.: Condens. Lorimer, Micheal Ostap, George Stan, Ryota Takaki, Ri- Matter 17 (47), S3851. Bhabha, G., H.-C. Cheng, N. Zhang, A. Moeller, M. Liao, ina Tehver, Huong Vu, Ahmet Yildiz, and Zhechun Zhang J. A. Speir, Y. Cheng, and R. D. Vale (2014), CELL for several useful discussions and collaborations involving 159 (4), 857. the topics covered in this review. DT is especially grate- Bhaskaran, H., and R. Russell (2007), Nature 449 (7165), ful to Prof. Michael E. Fisher for innumerable inspiring 1014. discussions for over twenty years not only on molecular Bianco, P. R., and S. C. Kowalczykowski (2000), Nature motors but also on the potential role of physics in biol- 405 (6784), 368. ogy. This work was supported in part by a grant from Bierbaum, V., and R. Lipowsky (2011), Biophysical journal 100 (7), 1747. National Science Foundation (CHE 19-00093) and the Block, S. M. (2007), Biophysical journal 92 (9), 2986. Collie-Welch Chair (F-0019) administered through the Block, S. M., C. L. Asbury, J. W. Shaevitz, and M. J. Lang Welch Foundation. (2003), Proceedings of the National Academy of Sciences 100 (5), 2351. Block, S. M., L. S. Goldstein, and B. J. Schnapp (1990), REFERENCES Nature 348 (6299), 348. Brady, S. T. (1985), Nature 317 (6032), 73. Alberts, B., A. Johnson, J. Lewis, M. Raff, K. Roberts, and Brown, A. I., and D. A. Sivak (2017), Proceedings of the P. Walter (2008), Molecular Biology of the Cell, 5th ed. National Academy of Sciences 114 (42), 11057. (Garland Science). Bustamante, C., W. Cheng, and Y. X. Mejia (2011), Cell Alhadeff, R., and A. Warshel (2017), Proc. Natl. Acad. Sci. 144 (4), 480. 114 (39), 10426. Callen, H. B. (1985), Thermodynamics and an Introduction Ali, J. A., and T. M. Lohman (1997), Science 275 (5298), to Thermostatistics (Wiley). 377. Caporizzo, M. A., C. E. Fishman, O. Sato, R. M. Jami- Altman, D., H. L. Sweeney, and J. A. Spudich (2004), Cell olkowski, M. Ikebe, and Y. E. Goldman (2018), Biophysical 116 (5), 737. journal 114 (6), 1400. Altschuler, S. J., and L. F. Wu (2010), Cell 141, 559. Carter, N. J., and R. Cross (2005), Nature 435 (7040), 308. Anderson, P. W. (1972), Science 177 (4047), 393. Chakrabarti, S., C. Hyeon, X. Ye, G. H. Lorimer, Andrecka, J., J. O. Arroyo, Y. Takagi, G. de Wit, A. Fineberg, and D. Thirumalai (2017), Proceedings of the National L. MacKinnon, G. Young, J. R. Sellers, and P. Kukura Academy of Sciences 114 (51), E10919. (2015), Elife 4, e05413. Chakrabarti, S., C. Jarzynski, and D. Thirumalai (2019), Ariga, T., M. Tomishige, and D. Mizuno (2018), Physical Biophysical Journal 10.1016/j.bpj.2019.07.021. review letters 121 (21), 218101. Cheng, W., S. Dumont, I. Tinoco, and C. Bustamante (2007), Astumian, R. D., and M. Bier (1996), Biophysical journal Proc. Natl. Acad. Sci. 104 (35). 70 (2), 637. Clancy, B. E., W. M. Behnke-Parks, J. O. Andreasson, S. S. Bagchi, D., M. Manosas, W. Zhang, K. A. Manthei, Rosenfeld, and S. M. Block (2011), Nature structural & S. Hodeib, B. Ducos, J. L. Keck, and V. Croquette (2018), molecular biology 18 (9), 1020. Nucleic acids research 46 (16), 8500. Clemen, A. E. M., M. Vilfan, J. Jaud, J. S. Zhang, M. Bar- Bagshaw, C., and D. Trentham (1973), Biochemical Journal mann, and M. Rief (2005), Biophys. J. 88 (6), 4402. 133 (2), 323. Comstock, M. J., K. D. Whitley, H. Jia, J. Sokoloski, T. M. Baker, J. E., E. B. Krementsova, G. G. Kennedy, A. Arm- Lohman, T. Ha, and Y. R. Chemla (2015), Science strong, K. M. Trybus, and D. M. Warshaw (2004), Pro- 348 (6232), 352. ceedings of the National Academy of Sciences 101 (15), Craig, E. M., and H. Linke (2009), Proceedings of the Na- 5542. tional Academy of Sciences 106 (43), 18261. 41

Crooks, G. E. (1998), Journal of Statistical Physics 90 (5-6), 248301. 1481. Goldtzvik, Y., M. Hinczewski, and D. Thirumalai (2019), De la Cruz, E. M., E. M. Ostap, and H. L. Sweeney (2001), preprint. Journal of Biological Chemistry 276 (34), 32373. Goldtzvik, Y., M. L. Mugnai, and D. Thirumalai (2018), De la Cruz, Enrique M, M., A. L. Wells, S. S. Rosenfeld, Structure 26 (12), 1664. E. M. Ostap, and H. L. Sweeney (1999), Proceedings of Golubeva, N., A. Imparato, and L. Peliti (2012), Euro Phys. the National Academy of Sciences 96 (24), 13726. Lett. 97 (6), 60005. De La Cruz, E. M., and E. M. Ostap (2004), Current opinion Gorbalenya, A. E., and E. V. Koonin (1993), Current opinion in cell biology 16 (1), 61. in structural biology 3 (3), 419. Dechant, A., and S.-i. Sasa (2018a), Journal of Statistical Gross, S. P., M. C. Tuma, S. W. Deacon, A. S. Serpinskaya, Mechanics: Theory and Experiment 2018 (6), 063209. A. R. Reilein, and V. I. Gelfand (2002), The Journal of Dechant, A., and S.-i. Sasa (2018b), Physical Review E cell biology 156 (5), 855. 97 (6), 062101. Guo, Z., and D. Thirumalai (1995), Biopolymers: Original Delagoutte, E., and P. H. Von Hippel (2002), Quarterly re- Research on Biomolecules 36 (1), 83. views of 35 (4), 431. Hackney, D. D. (1988), Proceedings of the National Academy Delagoutte, E., and P. H. Von Hippel (2003), Quarterly re- of Sciences 85 (17), 6314. views of biophysics 36 (1), 1. Hackney, D. D. (2005), Proceedings of the National Academy Derrida, B. (1983), Journal of statistical physics 31 (3), 433. of Sciences 102 (51), 18338. Dessinges, M.-N., T. Lionnet, X. G. Xi, D. Bensimon, and Hancock, W. O. (2016), Biophysical journal 110 (6), 1216. V. Croquette (2004), Proceedings of the National Academy Harada, T., and S.-i. Sasa (2006), Physical Review E 73 (2), of Sciences 101 (17), 6439. 026131. DeWitt, M. A., A. Y. Chang, P. A. Combs, and A. Yildiz Hartman, M. A., D. Finan, S. Sivaramakrishnan, and J. A. (2012), Science 335 (6065), 221. Spudich (2011), Ann. Rev. Cell DEv. Biol. 27, 133. Dominguez, R., Y. Freyzon, K. M. Trybus, and C. Cohen Hartman, M. A., and J. A. Spudich (2012), J. Cell. Sci. (1998), Cell 94 (5), 559. 125 (7), 1627. Dong, F., E. P. Gogol, and P. H. von Hippel (1995), Journal Hartwell, L. H., J. J. Hopfield, S. Leibler, and A. W. Murray of Biological Chemistry 270 (13), 7462. (1999), Nature 402 (6761supp), C47. Dunn, A. R., P. Chuan, Z. Bryant, and J. A. Spudich (2010), Hathcock, D., R. Tehver, M. Hinczewski, and D. Thirumalai Proceedings of the National Academy of Sciences 107 (17), (2019), . 7746. Hertzano, R., E. Shalit, A. K. Rzadzinska, A. A. Dror, Dunn, A. R., and J. A. Spudich (2007), Nature Structural L. Song, U. Ron, J. T. Tan, A. S. Shitrit, H. Fuchs, T. Has- and Molecular Biology 14 (3), 246. son, et al. (2008), PLoS genetics 4 (10), e1000207. Ehrenberg, M., and C. Blomberg (1980), Biophysical journal Hill, T. L. (1966), Journal of theoretical biology 10 (3), 442. 31 (3), 333. Hill, T. L. (2005a), Free Energy Transduction and Biochemical Elber, R., and A. West (2010), Proc. Natl. Acad. Sci. Cycle Kinetics (Dover). 107 (11), 5001. Hill, T. L. (2005b), Free energy transduction and biochemical Elting, M. W., Z. Bryant, J.-C. Liao, and J. A. Spudich cycle kinetics (Dover Publications, Inc.). (2011), Biophysical journal 100 (2), 430. Hill, T. L., and Y.-D. Chen (1975), Proceedings of the Na- Endow, S., and H. Higuchi (2000), Nature 406 (6798), 913. tional Academy of Sciences 72 (4), 1291. Fei, X., X. Ye, N. A. LaRonde, and G. H. Lorimer (2014), Hill, T. L., and R. Simmons (1976), Proceedings of the Na- Proceedings of the National Academy of Sciences 111 (35), tional Academy of Sciences 73 (1), 95. 12775. Hinczewski, M., R. Tehver, and D. Thirumalai (2013), Pro- Finer, J. T., R. M. Simmons, and J. A. Spudich (1994), ceedings of the National Academy of Sciences 110 (43), Nature 368 (6467), 113. E4059. Fisher, M. E., and A. B. Kolomeisky (1999a), Proceedings of Hirokawa, N., Y. Noda, Y. Tanaka, and S. Niwa (2009), Nat. the National Academy of Sciences 96 (12), 6597. Rev. Mol. Cell. Biol. 10 (10), 682. Fisher, M. E., and A. B. Kolomeisky (1999b), Physica A: Holzbaur, E. L., and Y. E. Goldman (2010), Current opinion Statistical Mechanics and its Applications 274 (1), 241. in cell biology 22 (1), 4. Fisher, M. E., and A. B. Kolomeisky (2001), Proceedings of Homma, K., and M. Ikebe (2005), Journal of Biological the National Academy of Sciences 98 (14), 7748. Chemistry 280 (32), 29381. Ge, H., and H. Qian (2010), Physical Review E 81 (5), Hopfield, J. J. (1974), Proceedings of the National Academy 051133. of Sciences 71 (10), 4135. Gebhardt, J. C. M., A. E.-M. Clemen, J. Jaud, and M. Rief Howard, J. (2001), Mechanics of motor proteins and the cy- (2006), Proceedings of the National Academy of Sciences toskeleton (Sinauer associates Sunderland, MA). 103 (23), 8680. Howard, J. (2006), Current Biology 16 (14), R517. Gennerich, A., A. P. Carter, S. L. Reck-Peterson, and R. D. Howard, J., A. Hudspeth, and R. Vale (1989), Nature Vale (2007), Cell 131 (5), 952. 342 (6246), 154. Gerritsma, E., and P. Gaspard (2010), Biophysical Reviews Hulett, H. (1970), Nature 225 (5239), 1248. and Letters 5 (04), 163. Hwang, W., and C. Hyeon (2017), The journal of physical Gibbons, I., and A. Rowe (1965), Science 149 (3682), 424. chemistry letters 8 (1), 250. Gingrich, T. R., J. M. Horowitz, N. Perunov, and J. L. Eng- Hwang, W., M. J. Lang, and M. Karplus (2008), Structure land (2016), Physical review letters 116 (12), 120601. 16 (1), 62. Gladrow, J., N. Fakhri, F. MacKintosh, C. Schmidt, and Hwang, W., M. J. Lang, and M. Karplus (2017), eLife 6, C. Broedersz (2016), Physical review letters 116 (24), e28948. 42

Hyeon, C., and W. Hwang (2017), Physical Review E 96 (1), Nucleic acids research 44 (9), 4330. 012156. Liepelt, S., and R. Lipowsky (2007a), Physical review letters Hyeon, C., S. Klumpp, and J. N. Onuchic (2009), Physical 98 (25), 258102. Chemistry Chemical Physics 11 (24), 4899. Liepelt, S., and R. Lipowsky (2007b), EPL (Europhysics Let- Hyeon, C., J. Lee, J. Yoon, S. Hohng, and D. Thirumalai ters) 77 (5), 50002. (2012), Nat. Chem. 4, 907. Lionnet, T., A. Dawid, S. Bigot, F.-X. Barre, O. A. Saleh, Hyeon, C., and J. N. Onuchic (2007a), Proceedings of the F. Heslot, J.-F. Allemand, D. Bensimon, and V. Croquette National Academy of Sciences 104 (7), 2175. (2006), Nucleic acids research 34 (15), 4232. Hyeon, C., and J. N. Onuchic (2007b), Proceedings of the Lipowsky, R., and S. Liepelt (2008), Journal of Statistical National Academy of Sciences 104 (44), 17382. Physics 130 (1), 39. Hyeon, C., and D. Thirumalai (2011), Nature communica- Lipowsky, R., S. Liepelt, and A. Valleriani (2009), Journal tions 2, 487. of Statistical Physics 135 (5), 951. Ikezaki, K., T. Komori, M. Sugawa, Y. Arai, S. Nishikawa, Liu, B., R. J. Baskin, and S. C. Kowalczykowski (2013), A. H. Iwane, and T. Yanagida (2012), Small 8 (19), 3035. Nature 500, 482. Ikezaki, K., T. Komori, and T. Yanagida (2013), PloS one Liu, C., X. Fu, L. Liu, X. Ren, C. K. Chau, S. Li, L. Xi- 8 (3), e58912. ang, H. Zeng, G. Chen, L.-H. Tang, et al. (2011), Science Inoue, A., J. Saito, R. Ikebe, and M. Ikebe (2002), Nature 334 (6053), 238. cell biology 4 (4), 302. Lohman, T. (1992), Molecular microbiology 6 (1), 5. Isojima, H., R. Iino, Y. Niitani, H. Noji, and M. Tomishige Lohman, T. M., and K. P. Bjornson (1996), Annual review (2016), Nature chemical biology 12 (4), 290. of biochemistry 65 (1), 169. Itoh, H., A. Takahashi, K. Adachi, H. Noji, R. Yasuda, Lohman, T. M., E. J. Tomko, and C. G. Wu (2008), Nature M. Yoshida, and K. Kinosita Jr (2004), Nature 427 (6973), Reviews Molecular Cell Biology 9 (5), 391. 465. Lucius, A. L., N. K. Maluf, C. J. Fischer, and T. M. Lohman Itsathitphaisarn, O., R. A. Wing, W. K. Eliason, J. Wang, (2003), Biophysical journal 85 (4), 2224. and T. A. Steitz (2012), Cell 151 (2), 267. Ma, J., P. Sigler, Z. Xu, and M. Karplus (2000), J. Mol. Biol. Iyer, L. M., D. D. Leipe, E. V. Koonin, and L. Aravind 302 (2), 303. (2004), Journal of structural biology 146 (1-2), 11. Ma, W., K. D. Whitley, Y. R. Chemla, Z. Luthey-Schulten, Jacob, F. (1977), Science 196 (4295), 1161. and K. Schulten (2018), eLife 7, e34186. Jeong, Y.-J., M. K. Levin, and S. S. Patel (2004), Proceedings Maes, C., and M. H. Van Wieren (2003), Journal of Statistical of the National Academy of Sciences 101 (19), 7264. Physics 112 (1-2), 329. Johnson, D. S., L. Bai, B. Y. Smith, S. S. Patel, and M. D. Manosas, M., X. G. Xi, D. Bensimon, and V. Croquette Wang (2007), Cell 129 (7), 1299. (2010), Nuc. Acids Res. 38 (16), 5518. J¨ulicher, F., A. Ajdari, and J. Prost (1997), Reviews of Mod- Marians, K. J. (2000), Structure 8 (12), R227. ern Physics 69 (4), 1269. Mehta, A. D., R. S. Rock, M. Rief, J. A. Spudich, M. S. Kad, N. M., K. M. Trybus, and D. M. Warshaw Mooseker, and R. E. Cheney (1999), Nature 400 (6744), (2008), Journal of Biological Chemistry 283 (25), 17477, 590. http://www.jbc.org/content/283/25/17477.full.pdf+html. Mehta, P., and D. J. Schwab (2012), Proceedings of the Na- Kiefhaber, T. (1995), Proceedings of the National Academy tional Academy of Sciences 109 (44), 17978. of Sciences 92 (20), 9029. Mickolajczyk, K. J., N. C. Deffenbaugh, J. O. Arroyo, J. An- Kirkpatrick, T., and D. Thirumalai (2015), Reviews of Mod- drecka, P. Kukura, and W. O. Hancock (2015), Proceed- ern Physics 87 (1), 183. ings of the National Academy of Sciences 112 (52), E7186. Kodera, N., D. Yamamoto, R. Ishikawa, and T. Ando (2010), Milic, B., J. O. Andreasson, D. W. Hogan, and S. M. Block Nature 468 (7320), 72. (2017), Proceedings of the National Academy of Sciences Kolomeisky, A. B., and M. E. Fisher (2000), Physica A: Sta- 114 (33), E6830. tistical Mechanics and its Applications 279 (1), 1. Mohr, S., J. M. Stryker, and A. M. Lambowitz (2002), Cell Kolomeisky, A. B., and M. E. Fisher (2003), Biophysical 109, 769. journal 84 (3), 1642. Moore, J. R., E. B. Krementsova, K. M. Trybus, and D. M. Kolomeisky, A. B., and M. E. Fisher (2007), Annu. Rev. Warshaw (2004), Journal of Muscle Research & Cell Motil- Phys. Chem. 58, 675. ity 25 (1), 29. Koza, Z. (1999), Journal of Physics A: Mathematical and Mugnai, M., M. Caporizzo, Y. Goldman, and D. Thirumalai General 32 (44), 7637. (2019), bioRxiv , 684696. Koza, Z. (2002), arXiv preprint cond-mat/0207203. Mugnai, M. L., and D. Thirumalai (2017), Proceedings of Lan, G., P. Sartori, S. Neumann, V. Sourjik, and Y. Tu the National Academy of Sciences 114 (22), E4389. (2012), Nature physics 8 (5), 422. Mukherjee, S., R. Alhadeff, and A. Warshel (2017), Proc. Lebowitz, J. L., and H. Spohn (1999), Journal of Statistical Natl. Acad. Sci. 114 (9), 2259. Physics 95 (1-2), 333. Mukherjee, S., and A. Warshel (2013), Proc. Natl. Acad. Sci. Leibler, S., and D. A. Huse (1993), The Journal of cell biology 110 (43), 17326. 121 (6), 1357. Nam, W., and B. I. Epureanu (2016), PloS one 11 (1), Levi, V., A. Serpinskaya, E. Gratton, and V. Gelfand (2006), e0147676. Biophys. J. 90 (1), 318. Nishikawa, S., I. Arimoto, K. Ikezaki, M. Sugawa, H. Ueno, Levin, M. K., and S. S. Patel (2002), Journal of Biological T. Komori, A. H. Iwane, and T. Yanagida (2010), Cell Chemistry 277 (33), 29377. 142 (6), 879. Li, J.-H., W.-X. Lin, B. Zhang, D.-G. Nong, H.-P. Ju, J.-B. Nishiyama, M., H. Higuchi, and T. Yanagida (2002), Nature Ma, C.-H. Xu, F.-F. Ye, X. G. Xi, M. Li, et al. (2016), Cell Biology 4 (10), 790. 43

Oguchi, Y., S. V. Mikhailenko, T. Ohki, A. O. Olivares, E. M. Rodenfels, J., K. M. Neugebauer, and J. Howard (2019), De la Cruz, and S. Ishiwata (2008), Proceedings of the Developmental cell 48 (5), 646. National Academy of Sciences 105 (22), 7714. Rosenfeld, S. S., and H. Lee Sweeney (2004), Jour- Oguchi, Y., S. V. Mikhailenko, T. Ohki, A. O. Olivares, E. M. nal of Biological Chemistry 279 (38), 40100, De La Cruz, and S. Ishiwata (2010), Nature chemical bi- http://www.jbc.org/content/279/38/40100.full.pdf+html. ology 6 (4), 300. Ross, J. L., K. Wallace, H. Shuman, Y. E. Goldman, and Oke, O. A., S. A. Burgess, E. Forgacs, P. J. Knight, E. L. F. Holzbaur (2006), Nat. Cell. Biol. 8 (6), 562. T. Sakamoto, J. R. Sellers, H. White, and J. Trinick (2010), Russell, R., I. Jarmoskaite, and A. M. Lambowitz (2013), Proceedings of the National Academy of Sciences 107 (6), RNA Biology 10 (1), 44. 2509. Sakamoto, T., I. Amitani, E. Yokota, and T. Ando (2000), Okten,¨ Z., L. S. Churchman, R. S. Rock, and J. A. Spudich Biochem. Biophys. Res. Commun. 272 (2), 586. (2004), Nature Structural and Molecular Biology 11 (9), Sakamoto, T., A. Yildiz, P. R. Selvin, and J. R. Sell- 884. ers (2005), Biochemistry 44 (49), 16203, pMID: 16331980, Oppenheim, I., K. E. Shuler, G. H. Weiss, et al. (1977), https://doi.org/10.1021/bi0512086. Stochastic processes in chemical physics (Mit Press). Sarlah, A., and A. Vilfan (2014), Biophys. J. 107 (3), 662. Ori-McKenney, K. M., J. Xu, S. P. Gross, and R. B. Vallee Sarlah,ˇ A., and A. Vilfan (2014), Biophysical journal 107 (3), (2010), Nature cell biology 12 (12), 1228. 662. Ortega, A., D. Amor´os, and J. G. de La Torre (2011), Bio- Schief, W. R., R. H. Clark, A. H. Crevenna, and J. Howard physical Journal 101 (4), 892. (2004), Proceedings of the National Academy of Sciences Pan, J., D. Thirumalai, and S. A. Woodson (1997), Journal 101 (5), 1183. of molecular biology 273 (1), 7. Schmidt, H., R. Zalyte, L. Urnavicius, and A. P. Carter Pang, P. S., E. Jankowsky, P. J. Planet, and A. M. Pyle (2015), Nature 518 (7539), 435. (2002), The EMBO journal 21 (5), 1168. Schmiedl, T., and U. Seifert (2008), EPL (Europhysics Let- Patel, S. S., and K. M. Picha (2000), Annual review of bio- ters) 83 (3), 30005. chemistry 69 (1), 651. Schnakenberg, J. (1976), Reviews of Modern physics 48 (4), Perkins, T. T., H.-W. Li, R. V. Dalal, J. Gelles, and S. M. 571. Block (2004), Biophysical journal 86 (3), 1640. Schnitzer, M. J., and S. Block (1995), in Cold spring har- Pierobon, P., S. Achouri, S. Courty, A. R. Dunn, J. A. Spu- bor symposia on quantitative biology, Vol. 60 (Cold Spring dich, M. Dahan, and G. Cappello (2009), Biophys. J. Harbor Laboratory Press) pp. 793–802. 96 (10), 4268. Schnitzer, M. J., and S. M. Block (1997), Nature 388 (6640), Pigolotti, S., I. Neri, E.´ Rold´an,and F. J¨ulicher (2017), Phys- 386. ical review letters 119 (14), 140604. Schnitzer, M. J., K. Visscher, and S. M. Block (2000), Nature Pincus, D. L., S. Chakrabarti, and D. Thirumalai (2015), cell biology 2 (10), 718. Biophysical Journal 109 (2), 220. Seifert, U. (2005a), Physical review letters 95 (4), 040602. Post, P. L., M. J. Tyska, C. B. O’Connell, K. Johung, A. Hay- Seifert, U. (2005b), EPL (Europhysics Letters) 70 (1), 36. ward, and M. S. Mooseker (2002), Journal of Biological Seifert, U. (2011a), Physical review letters 106 (2), 020601. Chemistry 277 (14), 11679. Seifert, U. (2011b), The European Physical Journal E 34 (3), Purcell, T. J., H. L. Sweeney, and J. A. Spudich (2005), 26. Proceedings of the National Academy of Sciences of the Seifert, U. (2012), Reports on progress in physics 75 (12), United States of America 102 (39), 13873. 126001. Pyle, A. M. (2008), Annu. Rev. Biophys. 37, 317. Sellers, J. (2000), Biochem. Biophys. ACTA - Mol. Cell Res. Pylypenko, O., L. Song, A. Shima, Z. Yang, A. M. Houdusse, 1496 (1), 3. and H. L. Sweeney (2015), Proc. Natl. Acad. Sci. 112 (11), Sellers, J. R., and C. Veigel (2010), Nature Structural and E1201. Molecular Biology 17 (5), 590. Qian, H. (2004), Physical Review E 69 (1), 012901. Shiraishi, N., K. Saito, and H. Tasaki (2016), Physical review Qian, H. (2007), Annu. Rev. Phys. Chem. 58, 113. letters 117 (19), 190601. Qian, H., and D. A. Beard (2005), Biophysical chemistry Soldati, T., and M. Schliwa (2006), Nature Reviews Molecu- 114 (2-3), 213. lar Cell Biology 7 (12), 897. Reck-Peterson, S. L., W. B. Redwine, R. D. Vale, and A. P. Solomatin, S. V., M. Greenfeld, S. Chu, and D. Herschlag Carter (2018), Nat. Rev. Mol. Cell. Biol. 19 (6), 382. (2010), Nature 463, 681. Reck-Peterson, S. L., A. Yildiz, A. P. Carter, A. Gennerich, Song, Y., J. O. Park, L. Tanner, Y. Nagano, J. D. Rabinowitz, N. Zhang, and R. D. Vale (2006a), Cell 126 (2), 335. and S. Y. Shvartsman (2019), Current Biology 29 (12), Reck-Peterson, S. L., A. Yildiz, A. P. Carter, A. Gennerich, R566. N. Zhang, and R. D. Vale (2006b), Cell 126 (2), 335. Spudich, J. A., and S. Sivaramakrishnan (2010), Nat. Rev. Rice, S., A. W. Lin, D. Safer, C. L. Hart, N. Naber, B. O. Mol. Cell. Biol. 11 (2), 128. Carragher, S. M. Cain, E. Pechatnikova, E. M. Wilson- Stan, G., B. Brooks, and D. Thirumalai (2005), J. Mol. Biol. Kubalek, M. Whittaker, et al. (1999), Nature 402 (6763), 350 (4), 817. 778. Sumi, T., and S. Klumpp (2019), Nano letters. Rocak, S., and P. Linder (2004), Nature reviews Molecular Sun, Y., and Y. E. Goldman (2011a), Biophys. J. 101 (1), 1. cell biology 5 (3), 232. Sun, Y., and Y. E. Goldman (2011b), Biophysical Journal Rock, R. S., S. E. Rice, A. L. Wells, T. J. Purcell, J. A. 101 (1), 1. Spudich, and H. L. Sweeney (2001), Proceedings of the Sun, Y., O. Sato, F. Ruhnow, M. E. Arsenault, M. Ikebe, National Academy of Sciences 98 (24), 13655. and Y. E. Goldman (2010), Nature structural & molecular biology 17 (4), 485. 44

Sweeney, H. L., and A. Houdusse (2010), Annual review of Veigel, C., S. Schmitz, F. Wang, and J. R. Sellers (2005), biophysics 39, 539. Nature cell biology 7 (9), 861. Tailleur, J., and M. Cates (2008), Physical review letters Veigel, C., F. Wang, M. L. Bartoo, J. R. Sellers, and J. E. 100 (21), 218103. Molloy (2002), Nature cell biology 4 (1), 59. Takaki, R., M. L. Mugnai, Y. Goldtzvik, and D. Thirumalai Velankar, S. S., P. Soultanas, M. S. Dillingham, H. S. Subra- (2019), in preparation. manya, and D. B. Wigley (1999), Cell 97 (1), 75. Tehver, R., and D. Thirumalai (2008), J. Mol. Biol. 377, Venkatesan, M., L. Silver, and N. Nossal (1982), Journal of 1279. Biological Chemistry 257 (20), 12426. Tehver, R., and D. Thirumalai (2010), Structure 18 (4), 471. Visscher, K., M. J. Schnitzer, and S. M. Block (1999), Nature Thirumalai, D., and B. Ha (1998), Theoretical and mathemat- 400, 184. ical models in polymer research, edited by A. Y. Grosberg Vu, H. T., S. Chakrabarti, M. Hinczewski, and D. Thirumalai (Academic Press, Boston). (2016), Physical review letters 117 (7), 078101. Thirumalai, D., and C. Hyeon (2005), Biochemistry 44 (13), Wagoner, J. A., and K. A. Dill (2016), The Journal of Phys- 4957. ical Chemistry B 120 (26), 6327. Thirumalai, D., C. Hyeon, P. I. Zhuravlev, and G. H. Lorimer Wagoner, J. A., and K. A. Dill (2019), Proceedings of the (2019), Chemical reviews 119 (12), 6788. National Academy of Sciences 116 (13), 5902. Thirumalai, D., and G. H. Lorimer (2001), Annual review of Walker, M. L., S. A. Burgess, J. R. Sellers, F. Wang, J. A. biophysics and biomolecular structure 30 (1), 245. Hammer III, J. Trinick, and P. J. Knight (2000), Nature Toba, S., T. M. Watanabe, L. Yamaguchi-Okimoto, Y. Y. 405 (6788), 804. Toyoshima, and H. Higuchi (2006), Proceedings of the Na- Walter, W. J., V. Ber´anek,E. Fischermeier, and S. Diez tional Academy of Sciences 103 (15), 5741. (2012), PLoS One 7 (8), e42218. Todd, M. J., G. H. Lorimer, and D. Thirumalai (1996), Proc. Wang, J., C. Herman, K. Tipton, C. Gross, and J. Weissman Natl. Acad. Sci. U. S. A. 93, 4030. (2002), Cell 111 (7), 1027. Tominaga, M., and A. Nakano (2012), Frontiers in plant sci- Wang, X., X. Zhang, D. Wu, Z. Huang, T. Hou, C. Jian, P. Yu, ence 3, 211. F. Lu, R. Zhang, T. Sun, et al. (2017), Elife 6, e23908. Toyabe, S., T. Okamoto, T. Watanabe-Nakayama, H. Take- Wong, I., and T. M. Lohman (1992), Science 256 (5055), tani, S. Kudo, and E. Muneyuki (2010), Physical review 350. letters 104 (19), 198103. Yildiz, A., J. N. Forkey, S. A. McKinney, T. Ha, Y. E. Gold- Toyabe, S., T. Watanabe-Nakayama, T. Okamoto, S. Kudo, man, and P. R. Selvin (2003), Science 300 (5628), 2061. and E. Muneyuki (2011), Proceedings of the National Yildiz, A., H. Park, D. Safer, Z. Yang, L.-Q. Chen, P. R. Academy of Sciences 108 (44), 17951. Selvin, and H. L. Sweeney (2004a), Journal of Biological Tsygankov, D., A. W. R. Serohijos, N. V. Dokholyan, and Chemistry 279 (36), 37223. T. C. Elston (2011), Biophys. J. 101 (1), 144. Yildiz, A., M. Tomishige, A. Gennerich, and R. D. Vale Turina, P., D. Samoray, and P. Gr¨aber (2003), The EMBO (2008), Cell 134 (6), 1030. Journal 22 (3), 418. Yildiz, A., M. Tomishige, R. D. Vale, and P. R. Selvin Uemura, S., H. Higuchi, A. O. Olivares, E. M. De La Cruz, (2004b), Science 303 (5658), 676. and S. Ishiwata (2004a), Nat. Struct. Mol. Biol. 11, 877. Yu, J., T. Ha, and K. Schulten (2006), Biophys. J. 91 (6), Uemura, S., H. Higuchi, A. O. Olivares, E. M. De La Cruz, 2097. and S. Ishiwata (2004b), Nat. Struct. Mol. Biol. 11 (9), Zhang, Y., and A. B. Kolomeisky (2018), J. Phys. Chem. B 877. 122 (13, SI), 3272. Vale, R. D. (2003a), Cell 112 (4), 467. Zhang, Z., Y. Goldtzvik, and D. Thirumalai (2017), Proceed- Vale, R. D. (2003b), J Cell Biol 163 (3), 445. ings of the National Academy of Sciences 114 (46), E9838. Vale, R. D., and R. A. Milligan (2000), Science 288 (5463), Zhang, Z., and D. Thirumalai (2012), Structure 20 (4), 628. 88. Zimmermann, E., and U. Seifert (2012), New Journal of Vale, R. D., T. S. Reese, and M. P. Sheetz (1985), Cell 42 (1), Physics 14 (10), 103023. 39. Van Kampen, N. (2007), Stochastic processes in physics and chemistry (Elsiver).