University of Pennsylvania ScholarlyCommons

Publicly Accessible Penn Dissertations

2014

Membrane Curvature Sorting And Generation By The Bar Domain Endophilin And Syndapin

Chen Zhu University of Pennsylvania, [email protected]

Follow this and additional works at: https://repository.upenn.edu/edissertations

Part of the Chemistry Commons

Recommended Citation Zhu, Chen, "Membrane Curvature Sorting And Generation By The Bar Domain Proteins Endophilin And Syndapin" (2014). Publicly Accessible Penn Dissertations. 1527. https://repository.upenn.edu/edissertations/1527

This paper is posted at ScholarlyCommons. https://repository.upenn.edu/edissertations/1527 For more information, please contact [email protected]. Membrane Curvature Sorting And Generation By The Bar Domain Proteins Endophilin And Syndapin

Abstract Membrane curvature provides a means to control spatial organization and activity of cells. It is regulated by plenty of peripherally binding membrane proteins, including BAR domain proteins. Two important sub- families of BAR domain containing proteins are NBAR and FBAR domain proteins. However, the current understanding of BAR domain membrane curvature (MC) sensing and generation is insufficient. My thesis intends to contribute to alleviating this situation.

We first performed curvature sorting and generation experiments of an NBAR domain protein: endophilin. We found that the endophilin NBAR domain (ENBAR) behaved as a curvature sensor or generator at different concentrations. We studied lateral diffusion of ENBAR and found its diffusion coefficients depending on its membrane density. We developed an analytical model to explain our experimental results. Our theory predicts that the membrane curvature serves as a switch that is modulated by a thermodynamic phase transition.

Then we studied the influence of membrane insertion helices on NBAR domain protein MC sensing and membrane dissociation kinetics by comparing ENBAR WT protein with helices deletion mutants. We found that the two helices had essential contributions for the NBAR domain curvature sorting. The WT and mutant membrane dissociation time were positively correlated with their densities on the membrane. Irreversible binding fractions for both variants were observed. These phenomena suggest higher order structure formation of these variants on the membrane.

Furthermore, we investigated the autoinhibition mechanism of full length endophilin via its H0 variants and the SH3 domain, by fluorescence experiments. We obtained evidence of the interaction between H0 helix and SH3 domain in solution and determined their binding affinities. Theseesults r experimentally support an H0/ SH3 domain mediated autoinhibition mechanism.

Finally we explored the regulation of a possible autoinhibition of the tubulation ability of syndapin, an FBAR domain protein. We compared the curvature sensing, generation and initiation abilities of syndapin FBAR, full length and full length with proline rich ligand. We found that the syndapin FBAR curvature initiation ability was larger compared to the full length protein, which was likely due to the differences in their curvature sensing abilities.

Degree Type Dissertation

Degree Name Doctor of Philosophy (PhD)

Graduate Group Chemistry

First Advisor Tobias Baumgart

Keywords BAR domain, biophysics, curvature, endocytosis, membrane, protein Subject Categories Chemistry

This dissertation is available at ScholarlyCommons: https://repository.upenn.edu/edissertations/1527 MEMBRANE CURVATURE SORTING AND GENERATION BY THE BAR DOMAIN

PROTEINS ENDOPHILIN AND SYNDAPIN

Chen Zhu A DISSERTATION in Chemistry Presented to the Faculties of the University of Pennsylvania in Partial Fulfillment of the Requirements for the Degree of Doctor of Philosophy 2014

Supervisor of Dissertation

______Dr. Tobias Baumgart Associate Professor of Chemistry

Graduate Group Chairperson

______

Dr. Gary A. Molander Hirschmann-Makineni Professor of Chemistry

Dissertation committee Dr. Jeffery G. Saven, Associate Professor of Chemistry Dr. Paul A. Janmey, Professor of Physiology Dr. Zahra Fakhraai, Assistant Professor of Chemistry

DEDICATION

To Haiying Zhu and Sui Xie

ii

ACKNOWLEDGEMENT The completion of this thesis is not possible without the help of many others. First of all,

I would like to take the opportunity to thank my thesis advisor Dr. Tobias Baugmart, who has always been very supportive since I joined the lab and served as an incredible advisor.

His patience and understanding helped me adapt the research at the initial phases of my research. His enthusiasm about science always inspired me. His depth and breadth of knowledge always impressed me. His guidance helped me throughout the experimental part of my PhD studies and writing of this thesis.

I am very grateful to have Dr. Jeffery G. Saven, Dr. Paul A. Janmey, Dr. Zahra Fakhraai as my committee members. I would like to thank Dr. Saven for serving as my committee chair. He has always been very kind, patient and gave me useful suggestions. Dr. Janmey shared his expertise and advice on membranes with me. I am thankful that Dr. Zahra

Fakhraai was willing to join my committee at the late stage of my research.

I thank the past and current lab members in Baumgart’s group. I learned a lot from several senior colleagues: Dr. Aiwei Tian, Dr. Benjamin R. Capraro and Dr. Michael C.

Heinrich helped me with tether pulling experiments, protein expression and purification, and the usage of optical tweezers. I also benefited from working with fellow graduate students: Tingting Wu, Zheng Shi, Zhiming Chen and other group members. It was a great pleasure and fortunate to work with so many excellent scientists.

Last but not the least, I would never have been able to finish this thesis without continuous supports and love from my friends and family, especially my parents, who are always there to cheer me up and stand by me.

iii

ABSTRACT MEMBRANE CURVATURE SORTING AND GENERATION BY THE BAR DOMAIN

PROTEINS ENDOPHILIN AND SYNDAPIN

Chen Zhu Dr. Tobias Baumgart

Membrane curvature provides a means to control spatial organization and activity of cells.

It is regulated by plenty of peripherally binding membrane proteins, including BAR domain proteins. Two important sub-families of BAR domain containing proteins are

NBAR and FBAR domain proteins. However, the current understanding of BAR domain protein membrane curvature (MC) sensing and generation is insufficient. My thesis intends to contribute to alleviating this situation.

We first performed curvature sorting and generation experiments of an NBAR domain protein: endophilin. We found that the endophilin NBAR domain (ENBAR) behaved as a curvature sensor or generator at different concentrations. We studied lateral diffusion of

ENBAR and found its diffusion coefficients depending on its membrane density. We developed an analytical model to explain our experimental results. Our theory predicts that the membrane curvature serves as a switch that is modulated by a thermodynamic phase transition.

Then we studied the influence of membrane insertion helices on NBAR domain protein

MC sensing and membrane dissociation kinetics by comparing ENBAR WT protein with helices deletion mutants. We found that the two helices had essential contributions for the

NBAR domain curvature sorting. The WT and mutant membrane dissociation time were positively correlated with their densities on the membrane. Irreversible binding fractions

iv for both variants were observed. These phenomena suggest higher order structure formation of these variants on the membrane.

Furthermore, we investigated the autoinhibition mechanism of full length endophilin via its H0 variants and the SH3 domain, by fluorescence experiments. We obtained evidence of the interaction between H0 helix and SH3 domain in solution and determined their binding affinities. These results experimentally support an H0/ SH3 domain mediated autoinhibition mechanism.

Finally we explored the regulation of a possible autoinhibition of the tubulation ability of syndapin, an FBAR domain protein. We compared the curvature sensing, generation and initiation abilities of syndapin FBAR, full length and full length with proline rich ligand.

We found that the syndapin FBAR curvature initiation ability was larger compared to the full length protein, which was likely due to the differences in their curvature sensing abilities.

v

TABLE OF CONTENTS

ACKNOWLEDGMENT ...... III

ABSTRACT ...... IV

TABLE OF CONTENTS ...... VI

LIST OF FIGURES AND TABLE ...... X

CHAPTER 1 Introdution ...... 1

A. Clathrin-mediated endocytosis ...... 1

B. BAR domain containing proteins ...... 2

1. MC S&G by BAR domain proteins ...... 3

2. Endophilin ...... 5

3. Syndapin ...... 7

C. References ...... 12

CHAPTER 2 Nonlinear sorting, curvature generation, and crowding of endophilin NBAR on tubular membranes ...... 17

A. Background ...... 17

B. Materials and methods ...... 18

1. Materials ...... 18

2. Methods...... 18

C. Results ...... 25

1. ENBAR curvature sorting on tubular membrane...... 25

2. ENBAR curvature sorting is in a concentration dependent manner ...... 26

3. ENBAR mobility on the tubular membrane is curvature dependent ...... 28

vi

D. Discussion ...... 30

1. Introduction of a nonlinear curvature/composition coupling model ...... 30

2. Comparison of analytical model to experimental data ...... 33

3. Possibility of curvature-induced phase transitions ...... 34

E. References ...... 47

CHAPTER 3 Membrane interacting helices influencing curvature sorting and membrane kinetics of endophilin NBAR ...... 49

A. Background ...... 49

B. Materials and methods ...... 50

1. Materials ...... 50

2. Methods...... 50

C. Results ...... 51

1. Membrane insertion amphipathic helices are essential for ENBAR curvature sensing ability ...... 51

2. Results from membrane dissociation kinetics of ENBAR WT and ΔH1I implied higher order oligomerization on the membrane ...... 53

D. Discussion ...... 54

E. References ...... 62

CHAPTER 4 Endophilin full length H0/SH3 intradimer / intermolecular autoinhibition ...... 64

A. Background ...... 64

B. Materials and methods ...... 65

1. Materials ...... 65

2. Methods...... 65

C. Results ...... 66

vii

1. H0F10FCN/ SH3 FRET experiments showed both donor and acceptor emission decreasing ...... 66

2. Trp in SH3 quenching induced by H0 F10W or H0 WT ...... 68

D. Discussion ...... 69

E. References ...... 78

CHAPTER 5 Syndapin full length BAR/ SH3 intramolecular autoinihbition ...... 80

A. Background ...... 80

B. Materials and methods ...... 81

1. Materials ...... 81

2. Methods...... 82

C. Results ...... 82

1. Curvature sorting of syndapin FBAR/ full length with or without the PRD domain ... 83

2. Curvature generation of syndapin FBAR/ full length with or without the PRD domain ...... 84

3. Curvature initiation of syndapin FBAR/ full length with or without the PRD domain 85

D. Discussion ...... 85

E. References ...... 92

CHAPTER 6 Conclusions and outlook ...... 93

A. Nonlinear sorting, curvature generation, and crowding of Endophilin NBAR on tubular membranes ...... 93

B. Membrane interacting helices influencing curvature sorting and membrane kinetics of endophilin NBAR ...... 93

C. Endophilin full length H0/ SH3 intradimer / intermolecular autoinhibition ...... 94

D. Syndapin full length BAR/ SH3 intramolecular autoinihibition ...... 95

E. References ...... 96

viii

APPENDIX ...... 97

BIBLIOGRAPHY ...... 100

ix

LIST OF FIGURES AND TABLE Table 1 Fit parameters for protein sorting and tether radius measurements 46

Figure 1.1 Curvature and clathrin-coated vesicle formation 9 Figure 1.2 Mechanisms of membrane curvature generation and sensing 10 Figure 1.3 Crystal structures of endophilin and syndapin 11 Figure 2.1 Rat endophilin A1 NBAR domains (ENBAR) partition in curvature gradients generated by tether membranes pulled from GUVs 37 Figure 2.2 ENBAR localization depends on membrane curvature 39 Figure 2.3 ENBAR diffusion on membrane tethers measured via fluorescence recovery after photobleaching (FRAP) slows with increasing curvature 41 Figure 2.4 Diffusion of ENBAR on membrane tethers is faster in tether elongation experiments compared to FRAP experiments at similar membrane tension 42 Figure 2.5 Comparison of curvature/ composition coupling model to experimental data 43 Figure 2.6 Van der Waals curvature/ composition coupling model predicts curvature- driven phase transition 44 Figure 3.1 Curvature preference partitioning of ENBAR WT, ENBAR ΔH1I and ENBAR ΔH0 57 Figure 3.2 Demonstration of ENBAR WT and mutants curvature sorting reversibilities 58 Figure 3.3 Comparison of curvature sensing abilities of ENBAR WT and mutants at the same protein membrane density 59 Figure 3.4 ENBAR ΔH1I dissociation follows single exponential decay 60 Figure 3.5 Monitoring ENBAR WT and ΔH1I membrane dissociation processes at different membrane densities 61 Figure 4.1 H0 F10FCN/ SH3 FRET 71 Figure 4.2 Determination of the linear regime of fluorescence intensity-concentration relationship for H0 F10FCN 73 Figure 4.3 Incubation of H0 F10FCN and SH3 leads to both donor and acceptor emission intensities decrease 74 Figure 4.4 Determine linear regime of fluorescence intensity-concentration relationship for endophilin SH3 domain 75 Figure 4.5 Titration of quenching of Trp in SH3 by H0 F10W or H0 WT 76 Figure 5.1 Curvature-dependent partition of syndapin FBAR (SFB) and syndapin full length (SFL) on liposome 87 Figure 5.2 Syndapin full length maintains curvature sorting ability, but weaker than syndapin FBAR 89 Figure 5.3 Syndapin full length could generate the spontaneous curvature of the membrane, as well as the syndapin FBAR 90 Figure 5.4 Syndapin full length is short of membrane curvature initiation ability 91

x

Chapter 1: Introductiona

A. Clathrin-mediated endocytosis (CME)

Clathrin-mediated endocytosis (CME) is a fundamental cellular process by which cells

internalize/ uptake molecules, especially nutrients(1).CME is essential for cellular signal

transduction and neurotransmission. It also serves as a controller to many eukaryotic

cellular membrane activities. At the early stage of CME, adaptor and receptor proteins

gather and nucleate on the plasma membrane, followed by clathrin coat assembly and

formation of a nascent clathrin coated pit. Then the curvature sensing/ generating proteins such as BAR domain proteins, ENTH and , gradually accumulate around the neck of the clathrin coated pit, stabilize it, squeeze the neck region, and finally finish the

scission/ release of the clathrin coated vesicle(2-3). Fig.1.1 is a canonical model for the

CME process.

The temporal and spatial organization of dozens of adaptor and accessory proteins

involved in CME is a field of active interest (4-6). Although the temporal and spatial organization of proteins involved in CME is important, the present work focuses on other

aspects in clathrin-mediated endocytosis. It addresses issues such as: how does the

membrane deform to different curvatures at different stages in CME? How do the above

mentioned adaptor and accessory proteins participate and contribute to finish this

aThis chapter is largely reproduced from previously published work: (1) Zhu, C., Das, S.L., and Baumgart,

T. (2012). Nonlinear sorting, curvature generation, and crowding of endophilin N-BAR on tubular membranes. Biophys. J.102, 1837-1845. (2) Baumgart, T., Capraro, B.R., Zhu, C. and Das, S.L. (2011).

Thermodynamics and mechanics of membrane curvature generation and sensing by proteins and .

Annu. Rev. Phys. Chem.62, 483-506.

1

essential cellular process? Here, we focus on the BAR domain proteins (described in

more detail in the following part) and investigate that how they sense and generate membrane curvature in CME in order to shed light on our understanding of the

mechanism of CME.

B. BAR domain containing proteins.

We are interested in studying the plasma membrane deformation by peripherally binding

membrane proteins involved in CME. We consider three major classes of proteins

involved in membrane curvature sensing and generation (MCS&G), roughly defined by

structural features. The first of these classes is represented by BAR domain family

proteins (Bin//Rvs). BAR domains are crescent-shaped dimeric α-helical

bundles that in many cases bind to membranes through both electrostatic and

hydrophobic interactions. Several different types of BAR domains can be distinguished

based on structural characteristics, including classical BAR, N-BAR, F-BAR, I-BAR, and

PX-BAR (7-9).

The second class is represented by dynamin family proteins; these proteins do not contain

BAR domains(10).

Proteins of those two families show structural features believed to generate MC via scaffolding (see below for a discussion of this mechanism).

The third class considered here includes proteins and protein domains not expected to exhibit scaffolding on the basis of structure. Rather, proteins belonging to the third class bear structural units that can lead to MCS&G by inserting (wedging) into the membrane.

Commonly, the inserting regions are intrinsically unfolded appendices that undergo 2

folding transitions to form amphipathic α-helices (AHs) upon membrane binding.

Members of this class include epsin N-terminal homology (ENTH) domain-containing proteins such as epsin, which is believed to be involved in clathrin-mediated endocytosis(11).

In this work we are mostly interested in the first class—BAR domain containing proteins.

This class of proteins is known for its membrane deformation ability, while the mechanisms for this capability are proposed to be a synergistic contribution from several following mechanisms:

1. MC S&G by BAR domain proteins.

1) Scaffolding (fig. 1.2A):

BAR domain–containing proteins and dynamin are believed to provide scaffolds for cylindrical curvature(12). Scaffolding is usually defined as an imprinting of protein monomer or oligomer intrinsic shape of the membrane-binding surface onto the underlying membrane. The proposal that scaffolding is responsible for MCS&G by BAR domains was prompted by the determination of the Drosophila amphiphysin N-BAR domain crystal structure(13). CryoEM reconstructions have supported this mechanism for the dynamin polymer(14-15).

2) Hydrophobic insertion (fig. 1.2B):

Hydrophobic insertion is another important mechanism proposed to contribute to membrane deformation induced by proteins. This mechanism is proposing that an amphipathic helix, which is usually disordered in solution, forms an alpha-helix and inserts into the bilayer during the membrane binding process, thus causing bilayer asymmetry and therefore generating membrane curvature. A representative example is 3

the ENTH domain(11). Membrane insertion of N-terminal AHs is also believed to be responsible for MC-S&G by Sar1(16), Arf1(17), and Arf6(18), as mutations in their AHs abolish tubulation. For BAR domain containing proteins, one subfamily are the NBAR domain proteins, which consist of an N-terminal amphipathic helix adjacent to the BAR domain. This N-terminal amphipathic helix folds upon membrane binding(13, 19-23)and is termed H0s. Mutational deletion of H0 from the amphiphysin 2 N-BAR domain reduced tubulation of vesicle membranes(13, 21).

3) Oligomerization (fig. 1.2C):

For lipids, quantitative experimental evidence(24-25) has shown, in combination with analytical thermodynamic/ mechanical models, that cooperativity can amplify curvature sorting(25). The association of F-BAR domain dimers into filaments leads to striated, lattice-like protein coats on lipid tubes (26-27). The efficacy of such amplification of MC sensing is further underscored by the dependence of dynamin polymerization on MC(28).

EM imaging(13, 29-31) and molecular dynamics simulations(32) suggest that endophilin and amphiphysin proteins and N-BAR domains also oligomerize on tubulated membranes.

H0s of these proteins may be involved in this oligomerization, as an amphiphysin H0 has been shown to form an anti-parallel spontaneous curvature: curvature of a membrane binding interface, or of a membrane itself, in the absence of any external stresses dimer in the membrane(20). H0-mediated oligomerization has also been suggested to contribute to

MCS&G for the non scaffolding ENTH domain(33).

4) Steric effect (fig. 1.2D):

It has previously been hypothesized(34-35), and more recently suggested by experiments(36), that local crowding of peripheral proteins can cause membrane bending. 4

In these experiments, only when local protein concentration was increased as a result of

lipid-based phase separation was tubulation observed. In this study, proteins were used

that are known to lack intrinsic curvature (i.e., scaffolding) effects, membrane-inserting

AHs, or a tendency for oligomerization(36). Bending by crowding therefore has to be

considered as a synergistic contributor to MCS&G.

2. Endophilin

Endophilin is an N-terminal BAR domain-containing protein(37-40) (crystal structure in fig. 1.3A&B), which is enriched at neural synapses. Endophilin assembles with dynamin and synaptojanin around the neck of clathrin-coated membrane invaginations(39, 41).

Endophilin also has been found to be involved in a clathrin-independent endocytic

pathway that is faster than clathrin-dependent endocytosis(42).

The Endophilin N-BAR domain (ENBAR) contains a BAR domain, an N-terminal helix

adjacent to the BAR domain (helix H0) and an additional amphipathic helix (H1 insert

helix, residues ~ 62-86)(19, 43-44). These amphipathic helices are disordered in aqueous

solution and form an α-helix upon membrane insertion(19, 45).

In vitro research has shown that endophilin senses MC and induces the deformation and tubulation of liposomes (19, 30, 44). The mechanism of MC generation and sensing by

endophilin is not fully understood. Liposome binding and tubulation assays, as well as

results based on electron paramagnetic resonance spectroscopy have suggested that the

concave surface of its BAR domain acts as a rigid, positively charged scaffold(19, 44),

which electrostatically interacts with negatively charged liposomes (12-13, 34). Electron

paramagnetic resonance spectroscopy measurements reported that the concave surface of 5

the endophilin BAR domain does not penetrate into the acyl chain level of the curved bilayer, implying that the BAR domain only peripherally interacts with the membrane(45). The rigidity and spontaneous curvature of the crescent shape are assumed to bend the membrane(46).

Interestingly, a recently developed single liposome membrane binding assay reported that the crescent shaped BAR domain dimer is not able to sense MC; instead, MC sensing was suggested to solely depend on the insertion of amphipathic helices into lipid packing defects(22). Indeed, H0 and the H1 insert helices are believed to drive MC (19, 44, 47), via their hydrophobic insertion into the membrane(12, 34, 45, 48). Furthermore, molecular dynamics simulations have shown that the H1 insert helix orients perpendicularly to the long axis of the N-BAR domains during membrane binding, and that the degree of membrane deformation is connected with H1 helix orientation(49).

Besides scaffolding and hydrophobic insertion, higher order oligomerization of BAR domain dimers may contribute to MC generation(34). Consistent with this hypothesis, striations have been observed on tubules generated via ENBAR domains(30). Theoretical characterization of the process of liposome tubulation(50) and vesiculation by N-BAR domains via mesoscopic simulations and electron microscopy imaging indicate an intricate coupling between protein density, degree of N-BAR oligomerization, and membrane deformation(31).

This dissertation sheds light on the understanding of the mechanism of cellular function of endophilin in CME process via in vitro approaches. For the NBAR domain of endophilin (ENBAR), we first experimentally characterized the effect of MC on both

ENBAR localization at different protein solution concentrations and translational 6

diffusion of ENBAR. We then derived an analytical curvature-sorting model that we compared to our data. Implications of this model for physiologically important membrane shape transitions are also discussed (Chapter 2).

We then aim to clarify the controversial contributions (see more information about this in

Chapter 3) of amphipathic helices in the MC S&G of endophilin in CME. We express and purify the endophilin NBAR H0-deletion mutant (ENBAR ΔH0) and H1I-deletion mutant (ENBAR ΔH1I), respectively. We compare the curvature sensing abilities of

ENBAR WT and their mutants. We find that at the similar protein membrane density level, both mutants’ curvature sorting efficiencies are lower than WT. This indicates an essential contribution of either the H0 helix or the H1I helix to the curvature sensing

ability of ENBAR. Likely, this amplifies the scaffolding effect induced by BAR domains.

We also monitor the influence of helices on the potential of ENBAR oligomerization, via

single GUV transfer experiments (Chapter 3).

Besides focusing on the NBAR domain of endophilin, we study how its SH3 domain

(ESH3) plays a role in CME. We observe the systematic quenching of fluorescence intensity of Trp when the H0 helix is titrated by the SH3 domain. We not only use H0

WT, but also H0-F10FCN and H0-F10W, as the quenchers, to monitor the intensity

decrease of Trp in the ESH3 domain. These findings suggest an interaction of endophilin

H0 and SH3, which a plausible mechanism for the autoinhibition of endophilin (Chapter

4).

3. Syndapin (pacsin)

Syndapin is an FBAR domain containing protein (crystal structure in fig. 1.3C&D). 7

Among three types of syndapin, syndapin I is enriched in neurons(51). Syndapin I is not only involved in clathrin-mediated endocytosis, but also interacts with dynamin to contribute to activity dependent bulk endocytosis (52). Besides, syndapin-N-WASP interaction could be an important coordinator for local actin polymerization(53).

Syndapin I consists of an FBAR domain, an NPF domain which mostly interacts with EH domain proteins, and an SH3 domain (7, 51, 53). The FBAR domain has been shown to

be able to induce different types of membrane deformations: tubes, pearling structures,

and small vesicles(54-56). This variety of shapes is likely induced by syndapin’s unique

“S” shaped structure (fig. 1.3D). There is a short loop within helix 2 which could function

as an amphipathic wedge inserted into the bilayer (57-58). This special hinge in pacsin 3

is also proposed to be the key coordinator for the radius of tubes it generated, and it may

regulate the higher order structure arrangement of pacsin 3(59).

The membrane tubulation ability of syndapin I full length is proposed to be inhibited in vivo(55). In vitro, full length syndapin I also tubulates membranes less efficiently compared to syndapin FBAR, as quantified by EM experiments (55-56). The current mechanism suggests an interaction between SH3 domain and FBAR domain in the full length protein, which inhibits the FBAR membrane bending ability. This mechanistic hypothesis calls for more for thorough analysis, as described in Chapter 5.

8

Figures:

Figure 1.1 Curvature and clathrin-coated vesicle formation. Diagram adapted from(1) with permission from Nature Publishing Group.

9

Figure 1.2 Mechanisms of membrane curvature generation and sensing. Figure is adapted from(1) with permission from Annual Reviews Publishing Group.

10

Figure 1.3 Crystal structures of endophilin and syndapin.

A&B. Ribbon diagram of rat endophilin-A1 BAR domain (Protein Data Bank (PDB)

accession number 2c08) from side view and top view.

C&D. Ribbon diagram of mouse syndapin I FBAR domain (Protein Data Bank (PDB)

accession number 2x3v) from side view and top view.

11

References:

1. McMahon, H. T., and E. Boucrot. 2011. Molecular mechanism and physiological functions of clathrin-mediated endocytosis. Nature Reviews Molecular Biology 12:517-533 2. Higgins, M. K., and H. T. McMahon. 2002. Snap-shots of clathrin-mediated endocytosis. Trends in Biochemical Sciences 27:257-263. 3. Doherty, G. J., and H. T. McMahon. 2009. Mechanisms of endocytosis. Annual Review of Biochemistry 78:857–902. 4. Taylor, M. J., D. Perrais, and C. J. Merrifield. 2011. A High Precision Survey of the Molecular Dynamics of Mammalian Clathrin-Mediated Endocytosis. Plos Biology 9:e1000604. 5. Cocucci, E., F. o. Aguet, S. Boulant, and T. Kirchhausen. 2012. The first five seconds in the life of a clathrin-coated pit. Cell 150:495–507. 6. Watanabe, S., B. R. Rost, M. Camacho-Pe´rez, M. W. Davis, B. So¨hl- Kielczynski, C. Rosenmund, and E. M. Jorgensen. 2013. Ultrafast endocytosis at mouse hippocampal synapses. Nature 504:242–247. 7. Frost, A., V. M. Unger, and P. De Camilli. 2009. The BAR domain superfamily: membrane-molding macromolecules. Cell 137:191-196. 8. Masuda, M., and N. Mochizuki. 2010. Structural characteristics of BAR domain superfamily to sculpt the membrane. Seminars in Cell & Developmental Biology 21: 391–398. 9. Prinz, W. A., and J. E. Hinshaw. 2009. Membrane-bending proteins. Critical Reviews in Biochemistry & Molecular Biology 44:278-291. 10. Praefcke, G. J. K., and H. T. McMahon. 2004. The dynamin superfamily: universal membrane tabulation and fission molecules? Nature Reviews Molecular Cell Biology 5:133-147. 11. Ford, M. G. J., I. G. Mills, B. J. Peter, Y. Vallis, G. J. K. Praefcke, P. R. Evans, and H. T. McMahon. 2002. Curvature of clathrin-coated pits driven by epsin. Nature 419:361-366. 12. Zimmerberg, J., and M. M. Kozlov. 2006. How proteins produce cellular membrane curvature. Nature Reviews Molecular Cell Biology 7:9-19. 13. Peter, B. J., H. M. Kent, I. G. Mills, Y. Vallis, P. J. G. Butler, P. R. Evans, and H. T. McMahon. 2004. BAR domains as sensors of membrane curvature: the amphiphysin BAR structure. Science 303:495-499. 14. Zhang, P., and J. E. Hinshaw. 2001. Three-dimensional reconstruction of dynamin in the constricted state. Nature Cell Biology 3:922 - 926. 15. Chen, Y.-J., P. Zhang, E. H. Egelman, and J. E. Hinshaw. 2004. The stalk region of dynamin drives the constriction of dynamin tubes. Nature Structural & Molecular Biology:574 - 575. 16. Lee, M. C. S., L. Orci, S. Hamamoto, E. Futai, M. Ravazzola, and R. Schekman. 2005. Sar1p N-terminal helix initiates membrane curvature and completes the fission of a COPII vesicle. Cell 122:605-617. 17. Krauss, M., J.-Y. Jia, A. Roux, R. Beck, F. T. Wieland, D. C. Pietro, and V. Haucke. 2008. Arf1-gtp-induced tubule formation suggests a function of arf 12

family proteins in curvature acquisition at sites of vesicle budding. The Journal of Biological Chemistry 283:27717-27723. 18. Lundmark, R., G. J. Doherty, Y. Vallis, B. J. Peter, and H. T. McMahon. 2008. Arf family GTP loading is activated by, and generates, positive membrane curvature. Biochemical Journal 414:189-194. 19. Gallop, J. L., C. C. Jao, H. M. Kent, P. J. G. Butler, P. R. Evans, R. Langen, and H. T. McMahon. 2006. Mechanism of endophilin N-BAR domain mediated membrane curvature. The EMBO Journal 25:2898-2910. 20. Fernandes, F., L. M. S. Loura, F. J. Chichon, J. L. Carrascosa, A. Fedorov, and M. Prieto. 2008. Role of helix 0 of the N-BAR domain in membrane curvature generation. Biophysical Journal 94:3065-3073. 21. Löw, C., U. Weininger, H. Lee, K. Schweimer, I. Neundorf, A. G. Beck-Sickinger, R. W. Pastor, and J. Balbach. 2008. Structure and dynamics of helix-0 of the N- BAR domain in lipid micelles and bilayers. Biophysical Journal 95:4315-4323. 22. Bhatia, V. K., K. L. Madsen, P. Bolinger, A. Kunding, P. Hedegard, U. Gether, and D. Stamou. 2009. Amphipathic motifs in BAR domains are essential for membrane curvature sensing. The EMBO Journal 28:3303-3314. 23. Hatzakis, N. S., V. K. Bhatia, J. Larsen, K. L. Madsen, P. Bolinger, A. H. Kunding, J. Castillo, U. H. Gether, P., and D. Stamou. 2009. How curved membranes recruit amphipathic helices and protein anchoring motifs. Nature Cell Biology 5:835-841. 24. Sorre, B., A. Callan-Jones, J. Manneville, P. Nassoy, J. Joanny, J. Prost, B. Goud, and P. Bassereau. 2009. Curvature-driven lipid sorting needs proximity to a demixing point and is aided by proteins. Proceedings of the National Academy of Science of the United States of America 106:5622-5626. 25. Tian, A., B. R. Capraro, C. Esposito, and T. Baumgart. 2009. Bending stiffness depends on curvature of ternary lipid mixture tubular membranes. Biophysical Journal 97:1636-1646. 26. Shimada, A., H. Niwa, K. Tsujita, S. Suetsugu, K. Nitta, K. Hanawa-Suetsugu, R. Akasaka, Y. Nishino, M. Toyama, L. Chen, Z. Liu, B. Wang, M. Yamamoto, T. Terada, A. Miyazawa, A. Tanaka, S. Sugano, M. Shirouzu, K. Nagayama, T. Takenawa, and S. Yokoyama. 2007. Curved EFC/F-BAR-domain dimers are joined end to end into a filament for membrane invagination in endocytosis. Cell 129:761-772. 27. Frost, A., R. Perera, A. Roux, K. Spasov, O. Destaing, E. H. Egelman, P. De Camilli, and V. M. Unger. 2008. Structural Basis of Membrane Invagination by F-BAR Domains. Cell 132:807-817. 28. Roux, A., G. Koster, M. Lenz, B. Sorre, J.-B. Manneville, P. Nassoy, and P. Bassereaua. 2010. Membrane curvature controls dynamin polymerization. Proceedings of the National Academy of Science of the United States of America 107:4141-4146. 29. Takei, K., V. I. Slepnev, V. Haucke, and P. D. Camilli. 1999. Functional partnership between amphiphysin and dynamin in clathrinmediated endocytosis. Nature Cell Biology 1:33-39.

13

30. Farsad, K., N. Ringstad, K. Takei, S. R. Floyd, K. Rose, and P. D. Camilli. 2001. Generation of high curvature membranes mediated by direct endophilin bilayer interactions. The Journal of Cell Biology 155:193-200. 31. Mizuno, N., C. C. Jao, R. R. Langen, and A. C. Steven. 2010. Multiple modes of endophilin-mediated conversion of lipid vesicles into coated tubes. The Journal of Biological Chemistry 285:23351-23358. 32. Ayton, G. S., and G. A. Voth. 2010. Multiscale simulation of protein mediated membrane remodeling. Seminars in Cell & Developmental Biology 21:357-362. 33. Yoon, Y., J. Tong, P. J. Lee, A. Albanese, N. Bhardwaj, M. Källberg, M. A. Digman, H. Lu, E. Gratton, Y.-K. Shin, and W. Cho. 2010. Molecular basis of the potent membrane-remodeling activity of the epsin 1 N-terminal homology domain. The Journal of Biological Chemistry 285:531-540. 34. McMahon, H. T., and J. L. Gallop. 2005. Membrane curvature and mechanisms of dynamic remodelling. Nature 438:590-596. 35. Sens, P., and M. S. Turner. 2004. Theoretical Model for the Formation of Caveolae and Similar Membrane Invaginations. Biophysical Journal 86:2049- 2057. 36. Stachowiak, J. C., C. C. Hayden, and D. Y. Sasaki. 2010. Steric confinement of proteins on lipid membranes can drive curvature and tubulation. Proceedings of the National Academy of Science of the United States of America 107:7781–7786. 37. Ringstad, N., H. Gad, P. Löw, G. Di Paolo, L. Brodin, O. Shupliakov, and P. De Camilli. 1999. Endophilin/SH3p4 is required for the transition from early to late stages in clathrin-mediated synaptic vesicle endocytosis. Neuron 24:143-154. 38. Verstreken, P., O. Kjaerulff, T. E. Lloyd, R. Atkinson, Y. Zhou, I. A. Meinertzhagen, and H. J. Bellen. 2002. Endophilin mutations block clathrin- mediated endocytosis but not neurotransmitter release. Cell 109:101-112. 39. Schuske, K. R., J. E. Richmond, D. S. Matthies, W. S. Davis, S. Runz, D. A. Rube, A. M. van der Bliek, and E. M. Jorgensen. 2003. Endophilin is required for synaptic vesicle endocytosis by localizing synaptojanin. Neuron 40:749-762. 40. Dickman, D. K., J. A. Horne, I. A. Meinertzhagen, and T. L. Schwarz. 2005. A slowed classical pathway rather than kiss-and-run mediates endocytosis at synapses lacking synaptojanin and endophilin. Cell 123:521-533. 41. Sundborger, A., C. Soderblom, O. Vorontsova, E. Evergren, J. E. Hinshaw, and O. Shupliakov. 2011. An endophilin-dynamin complex promotes budding of clathrin-coated vesicles during synaptic vesicle recycling. Journal of Cell Science 124:133-143. 42. Llobet, A., J. L. Gallop, J. J. E. Burden, G. Camdere, P. Chandra, Y. Vallis, C. R. Hopkins, L. Lagnado, and H. T. McMahon. 2011. Endophilin drives the fast mode of vesicle retrieval in a ribbon synapse. The Journal of Neuroscience 31:8512- 8519. 43. Weissenhorn, W. 2005. Crystal structure of the Endophilin-A1 BAR domain. Journal of Molecular Biology 35:653-661. 44. Masuda, M., S. Takeda, M. Sone, T. Ohki, H. Mori, Y. Kamioka, and N. Mochizuki. 2006. Endophilin BAR domain drives membrane curvature by two newly identified structure-based mechanisms. The EMBO Journal 25:2889-2897. 14

45. Jao, C. C., B. G. Hegde, J. L. Gallop, P. B. Hegde, H. T. McMahon, I. S. Haworth, and R. Langen. 2010. Roles of Amphipathic Helices and the Bin/Amphiphysin/Rvs (BAR) Domain of Endophilin in Membrane Curvature Generation. The Journal of Biological Chemistry 285:20164-20170. 46. Zimmerberg, J. 2004. Membrane curvature: How BAR domains bend bilayers Curr. Biol. 14:R250-R252. 47. Campelo, F., G. Fabrikant, H. T. McMahon, and M. M. Kozlov. 2010. Modeling membrane shaping by proteins: focus on EHD2 and N-BAR domains. FEBS Letters 584:1830-1839. 48. Campelo, F., H. T. McMahon, and M. M. Kozlov. 2008. The hydrophobic insertion mechanism of membrane curvature generation by proteins. Biophysical journal 95:2325-2339. 49. Cui, H., G. S. Ayton, and G. A. Voth. 2009. Membrane binding by the endophilin N-BAR domain. Biophysical Journal 97:2746-2753. 50. Ayton, G. S., E. Lyman, V. Krishna, R. D. Swenson, C. Mim, V. M. Unger, and G. A. Voth. 2009. New insights into BAR domain-induced membrane remodeling. Biophysical journal 97:1616-1625. 51. Modregger, J., B. Ritter, B. Witter, M. Paulsson, and M. Plomann. 2000. All three PACSIN isoforms bind to endocytic proteins and inhibit endocytosis. Journal of Cell Science 113:4511-4521. 52. Clayton, E. L., V. Anggono, K. J. Smillie, N. Chau, P. J. Robinson, and M. A. Cousin. 2009. The phospho-dependent dynamin-syndapin interaction triggers activity-dependent bulk endocytosis of synaptic vesicles. The Journal of Neuroscience 29:7706–7717. 53. Kessels, M. M., and B. Qualmann. 2002. Syndapins integrate N-WASP in receptor-mediated endocytosis. The EMBO Journal 21:6083-6094. 54. Wang, Q., M. V. A. S. Navarro, G. Peng, E. Molinelli, S. Lin Goh, B. L. Judson, K. R. Rajashankar, and H. Sondermann. 2009. Molecular mechanism of membrane constriction and tubulation mediated by the F-BAR protein Pacsin/Syndapin. Proceedings of the National Academy of Science of the United States of America 106:12700-12705. 55. Rao, Y., Q. Ma, A. Vahedi-Faridi, A. Sundborger, A. Pechstein, D. Puchkov, L. Luo, O. Shupliakov, W. Saenger, and V. Haucke. 2010. Molecular basis for SH3 domain regulation of F-BAR-mediated membrane deformation. Proceedings of the National Academy of Science of the United States of America 107:8213-8218. 56. Goh, S. L., Q. Wang, L. J. Byrnes, and H. Sondermann. 2012. Versatile Membrane Deformation Potential of Activated Pacsin. Plos ONE 7:e51628. 57. Edeling, M. A., S. Sanker, T. Shima, P. K. Umasankar, S. Höning, H. Y. Kim, L. A. Davidson, S. C. Watkins, M. Tsang, D. J. Owen, and L. M. Traub. 2009. Structural requirements for PACSIN/Syndapin operation during zebrafish embryonic notochord development. Plos ONE 4:e8150. 58. Plomann, M., J. G. Wittmann, and M. G. Rudolph. 2010. A hinge in the distal end of the PACSIN 2 F-BAR domain may contribute to membrane-curvature sensing. Journal of Molecular Biology 400:129–136.

15

59. Bai, X., G. Meng, M. Luo, and X. Zheng. 2012. Rigidity of wedge loop in PACSIN 3 protein Is a key factor in dictating diameters of tubules. The Journal of Biological Chemistry 287:22387-22396.

16

Chapter 2: Nonlinear sorting, curvature generation, and crowding of Endophilin

NBAR on tubular membranesa

A. Background

The curvature of biological membranes is controlled by membrane-bound proteins. For

example, during endocytosis, the sorting of membrane components, vesicle budding, and

fission from the plasma membrane are mediated by adaptor and accessory proteins.

Endophilin is a peripherally binding membrane protein that functions as an endocytic

accessory protein.

Endophilin’s membrane tubulation capacity is well-known. However, in order to understand thermodynamic and mechanical aspects of endophilin function, experimental measurements need to be compared to quantitative theoretical models.

We present measurements of curvature sorting and curvature generation of the endophilin

A1 NBAR domain on tubular membranes pulled from giant unilamellar vesicles (GUV).

At low concentration, endophilin functions primarily as a membrane curvature sensor; at

high concentrations, it additionally generates curvature. We determine the spontaneous

curvature induced by endophilin, and observe sigmoidal curvature/composition coupling

isotherms that saturate at high membrane tensions and protein solution concentrations.

The observation of saturation is supported by a strong dependence of lateral diffusion

coefficients on protein density on the tether membrane.

a This chapter is largely reproduced from previously published work: Zhu, C., Das, S.L., and Baumgart, T.

(2012). Nonlinear sorting, curvature generation, and crowding of endophilin NBAR on tubular membranes.

Biophys. J. 102, 1837-1845.

17

We develop a non-linear curvature/composition coupling model that captures our

experimental observations. Our model predicts a curvature-induced phase transition

among two states with varying protein density and membrane curvature. This transition

could act as a switch during endocytosis.

B. Materials and methods

1. Materials 1,2-dioleoyl-sn-glycero-3-phosphocholine (DOPC), 1,2-dioleoyl-sn- glycero-3-phospho-(1’-rac-glycerol) (DOPG) and 1,2-distearoyl-sn-glycero-3-

phosphoethanolamine-N-[biotinyl(polyethylene glycol)-2000] (ammonium salt) (DSPE-

Bio-PEG2000) were obtained from Avanti Polar Lipids (Alabaster, AL). Fatty-acid free

bovine serum albumin (BSA) was from Sigma Chemical (St. Louis, MO). Rat endophilin

A1 NBAR-Alexa Fluor 488 (ENBAR-A488, amino acids 1-247, labeled at C108) was

obtained from Pr. Langen, U. Southern California, USA, and stored in buffer (20 mM

HEPES, pH 7.4, 150 mM NaCl). Using vesicle spin down assays(1), we confirmed that

fluorescence labeling at position C108 does not alter membrane binding (data not shown).

Texas Red-1,2-dihexadecanoyl-sn-glycero-3-phosphoethanolamine triethylammonium

salt (TR-DHPE) was from Invitrogen (Carlsbad, CA). Streptavidin conjugated

microspheres with diameter of 6 μm were from Polysciences, Inc. (Warrington, PA).

2. Methods:

Preparation of giant unilamellar vesicles (GUVs): GUVs were prepared by

electroformation in solutions of 300 mM sucrose as described(2). Lipids were mixed in

18

chloroform at a total concentration of 1 mM. DOPG was used at 25 mol%, DOPC at 74

mol%, TR-DHPE at 0.3 mol%, and DSPE-Bio-PEG2000 content was 0.7 mol%.

ENBAR-A488 was added after electroswelling (so proteins bind to the exterior leaflet of

GUV membrane only) but immediately before micropipette aspiration experiments, to yield final solution concentrations indicated below.

Preparation of micropipettes: Micropipettes (World Precision Instruments, Sarasota, FL)

were manufactured with a pipette puller; pipette tips were clipped using a microforge.

The inner opening diameters were 3-6 μm. Irreversible membrane/pipette adhesion was

avoided by incubating micropipette tips with 0.5 mg/mL fatty-acid-free BSA dissolved in

1X PBS with a MicroFil needle (WPI), followed by rinsing, and pipettes were finally filled with 300 mM sucrose solution.

Preparation of chamber and tether pulling: GUV dispersions obtained through

electroswelling were diluted 1:10 in 300 mM sucrose solution. 50 μL diluted GUV

dispersions which contained ~0.5 μL 10X PBS, 0.5-1 μL streptavidin coated polystyrene

bead solution and various concentrations of ENBAR-A488 (rat endophilin A1 NBAR,

amino acids 1-247, labeled at C108 with AlexaFluor 488) solution were injected into a

measurement chamber. The chamber was constructed from microscope slides and

coverslips that allowed access by two perpendicularly oriented micropipettes.

Micropipettes were inserted into the chamber by a three-dimensional motorized manipulator system (Luigs and Neumann, Ratingen, Germany). Such a vesicle was

19

pipette-aspirated with an initial suction pressure amounting to 5~10 Pa. The aspiration pressure was controlled by adjusting the water level of a reservoir connected to the micropipettes, and monitored by a pressure transducer with a DP-28 diaphragm

(Validyne Engineering, Los Angeles, CA). In order to pull tethers, a second pipette was used to aspirate a bead. The bead was gently moved toward the aspirated vesicle and contacted for ~1 min, and then moved away from the vesicle to pull a membrane tether of

5-15 μm in length, depending on vesicle size and excess area. All experiments were carried out at room temperature (20 ± 2 °C).

Imaging: Vesicles and tethers were imaged with a fluorescence confocal microscopy

(FV3000 scanning system integrated with a motorized inverted microscope IX81;

Olympus, Center Valley, PA), using a 60X, 1.2 NA water immersion lens (Olympus).

Image analysis was carried out via IMAGEJ (National Institutes of Health, Bethesda,

MD). Vesicle fluorescence intensity values were measured after background subtraction from the average of four randomly chosen equal area regions of interest on the vesicle equator.

Tether cross section fluorescence intensity measurements: To investigate protein partitioning driven by MC, we monitored the local fluorescence intensities on the tubular membrane under varying membrane tensions. We changed membrane tension by adjusting the height of a water reservoir. Fluorescence intensities of tethers were measured by obtaining Kalman-averaged images of the tether cross section (xz plane),

20

which is orthogonal to the axis of the tether (contained in the xy plane; Fig. 2.1A), at a

stepwidth of 0.15 μm to yield a total imaging depth of around 6 μm. Cross-sectional fluorescence intensity profiles (Fig. 2.1B) were background-corrected, and intensity was

evaluated in an elliptical region of interest.

In order for us to be able to correlate membrane tether fluorescence intensity changes

with changes in protein coverage fraction, we determined the linear range of fluorescence.

For this purpose, tethers were pulled from pipette-aspirated GUVs incubated with

ENBAR-A488 at a concentration of 150 nM. Intensities of tethers at fixed membrane tension were measured for varying laser powers(3). Consequently, all measurements in this report were obtained using laser powers within the indicated linear range.

Determination of dissociation constant for ENBAR-A488 binding on GUVs: Vesicles were prepared as described above. GUV membranes contained 74% DOPC, 25% DOPG,

0.3% Texas Red-DHPE and 0.7% DSPE-Bio-PEG2000, and were diluted 1:10 in 300

mM sucrose. Protein solutions were mixed with vesicles and incubated for 30 min at

room temperature before imaging. Vesicle fluorescence intensities were determined as

described above. Vesicle fluorescence intensity values as a function of protein solution

concentration were fitted by a classical Langmuir adsorption isotherm.

Diffusion measurements on tethers via fluorescence recovery after photobleaching: For

fluorescence recovery after photobleaching (FRAP) measurements on the membrane

21

tether, tethers with lengths of 12 ± 1 μm were pulled and kept at fixed membrane tension.

Except for a short stretch amounting to a length of ~0.5 μm measured from the

tether/vesicle junctions, the entire tether was photobleached by repeated scanning at

maximal laser intensity (488 nm illumination). Prebleach and postbleach intensities were

measured using excitation with small laser power (0.1%~0.3% of full power) at 488 nm.

The relative photobleaching recovery ratio R(t) at a given time t was defined as

I(t) − I(0) R(t) = (1) I(−) − I(0)

where I(t), I(0)and I( ) are the fluorescence intensities of the tether integrated along

length increments dx at− time t, at a time immediately after bleaching (t = 0), and before

bleaching, respectively. A one-dimensional diffusion model(4) was fit to the

photobleaching recovery ratio R(t)

1 L2 1 h − x + L h + x − L R(t) = dxρ0{1− [erf ( ) + erf ( )]} (2) ∫L L 1 2 2 Dt 2 Dt

where D is the diffusion coefficient, ρ0 is the maximal local recovery ratio, which is

positive and no larger than 1. L1 and L2 are the beginning and ending position of the

analysis range on the tether. L is total length of the region of interest, and h is the position of an image source accounting for the presence of an impermeable boundary at the bead position(4). Fitting was done via the software Mathematica (Wolfram Research,

Champaign, IL).

Measurement of tube radius: The radii of membrane tubes displayed in Figs. 2.2E and

2.2F in the main text is below the resolution of the optical microscope. It is possible to

22

estimate it after calibrating the fluorescence intensity measurements of tubes and vesicles

via fluorescently labeled lipid(5). The tube radius (Rt) is proportional to the fluorescence

lipid intensity from the lipid dye on the tube (It ) normalized by the same intensity on the

lipid vesicle (Iv ).

I lipid R = A⋅ t t I lipid V (3)

We experimentally determined the calibration factor A from a linear fit to a plot of the theoretically expected tube radius (varied through membrane tension, in the absence of protein) against this ratio, yielding a slope of A = 229 ± 40 nm (uncertainty is the standard deviation of three independent experiments). We then used this conversion factor to extract the tube radius from the curvature sorting experiments.

Error estimation of uncertainty in diffusion coefficient: We use the χ–square test (6) to

calculate uncertainties in diffusion coefficients. To calculate χ2, we varied diffusion

coefficients while fixing the fitting parameters ρ0, as well as L1 and L2 (see main text for

definitions of these parameters) and summed residuals according to

n data fit 2 [R(t j ) − R(t j ) ] χ(D (Σ))2 = Di i ∑ 2 (4) j=1 σ

where σ is the relative uncertainty of the fluorescence intensity of an unbleached tether determined by image analysis, and Di are values for the diffusion coefficient close to the

optimal fit value. By calculating the following 2nd order differential, we obtained

uncertainties s for the diffusion coefficients (see Fig. 2.3 D):

23

∂ 2 χ(D )2 = i s 2/( 2 ) (5) ∂Di

Error estimation of uncertainty in fitting parameters for VdW model: The uncertainties

for fit parameters were determined by χ–square test. To obtain the uncertainty of the fit parameter ai (i = 1-4 for a, b, θves and Cp, respectively), we held ak(k� i) constant, and

fit varied aii around ai ,

n [R(Σ )data − R(Σ ) fit ]2 χ 2 = j j aii (aii ) ∑ 2 (6) j=1 σ

where σ is the relative uncertainty of the fluorescence intensity as defined above.

nd 2 Calculating the 2 order differential of χ with respect to the parameter aii gives the

uncertainty si for the fit parameter ai:

2 2 ∂ χ(aii ) si = 2 /( 2 ) (7) ∂aii

Note that this approach neglects covariant terms in the error matrix.

Error propagation for membrane tension: To estimate the uncertainty of membrane

tension Σ (see Figs. 2.2 C-F), we used multivariate error analysis(6-7):

δΣ 2 δ (∆P) 2 Rv 2 δRp 2 Rp 2 δRv 2 ( ) = ( ) + ( ) ( ) + ( ) ( ) (8) Σ ∆P Rv − Rp Rp Rv − Rp Rv

The uncertainty of aspiration pressure (δ(ΔP)) was 0.5 Pa. Since the uncertainty in the pipette radius contributes a constant error to membrane tension(8), we ignored this

24

component in our error analysis. The error for the vesicle radius (δRv ) was approximated as the image resolution, which is around 0.25 µm.

C. Results.

1. ENBAR curvature sorting on tubular membrane

An important goal of this study was to investigate to what extent and under which

conditions the NBAR domain of the peripheral protein endophilin A1 (ENBAR, from rat)

binds to tubular membranes with variable curvature. Electron microscopy observations

(9-11) combined with computational studies (12-14) have already demonstrated that

ENBAR can deform membranes into high curvature assemblies (with varying curvature

radii typically significantly below 50 nm at multi-micromolar protein concentrations). To be able to understand thermodynamic and mechanical aspects of ENBAR domain function, however, requires a comparison of measurements to quantitative models. In

order to quantitatively characterize the curvature sensing of ENBAR, we incubated

ENBAR-A488 with negatively-charged GUVs containing the lipid fluorophore TR-

DHPE, and pulled cylindrical tethers from pipette-aspirated vesicles, using streptavidin- conjugated microspheres. Fig. 2.1A demonstrates qualitatively that green (ENBAR protein) fluorescence is enriched on highly curved tubular membranes, rather than on the quasi-flat vesicular membrane (partially shown on the right-hand side in the fluorescence micrographs of Fig. 2.1A).

The curvature-induced partitioning of ENBAR was determined by confocal microscopy fluorescence imaging of membrane tether cross sections (Fig. 2.1B) and analyzed as described in the Materials and Methods section. Fig. 2.1C shows that green fluorescence 25

intensity (ENBAR) increases on the tether as membrane tension is increased, whereas the

lipid membrane tether fluorescence decreases as a consequence of the shrinking tubular

radius(15).

Thermodynamic interpretation of our data (see below) requires assessment of

reversibility and equilibration times of fluorescence intensities of the protein and the lipid

probe under varying membrane tension. Both green and red fluorescence signals respond

to large rapid (roughly 0.5mN/m/min) increases of membrane tension within about 1 min

and reach equilibrium (see Fig. 2.1C and fluorescence intensity ratio shown in Fig. 2.1D).

Subsequently lowering tension causes corresponding fluorescence intensity changes,

which demonstrates reversibility of the measurements.

2. ENBAR curvature sorting is in a concentration dependent manner

In Fig. 2.2A we display the analysis of a typical ENBAR curvature sorting experiment

using a protein solution concentration of 40 nM. With increasing lateral tension,

fluorescence intensity in the green (protein) fluorescence channel monotonically

increases, whereas the opposite is observed in the red (lipid) channel. In Fig. 2.2,

fluorescence intensity measurements are plotted against the square root of lateral tension

for the following reason. For the case of linear curvature sorting, the square root of lateral

tension can be shown to be proportional to membrane curvature (16). The plots in Fig.

2.2 therefore allow assessment of the linearity of curvature sorting. Importantly, the

results shown here demonstrate non-linear curvature/composition coupling; hence they

deviate from those found for the epsin N-terminal homology (ENTH) domain, where

sorting was observed to be proportional to the square root of membrane tension(17). 26

The fluorescence intensity of lipid probes in high curvature tether membranes used in this

work is linearly proportional to the MC (see supplementary materials of (3), consistent

with our previous findings(8, 15)); fluorescent lipids therefore are not significantly sorted

by membrane curvature and here serve as a reference for ratiometric fluorescence

intensity measurements. Fig. 2.2B shows the ratio Ir of protein and lipid probe

fluorescence intensities (Ir= Igreen / Ired) for the same data as shown in Fig. 2.2A.

To facilitate the comparison of our data to a thermodynamic model (see below), the

0 0 relative fluorescence intensities Ir were normalized to values I r ( I r = Ives-green / Ives-red) measured on the vesicle (described in Materials and Methods). A series of individual

sorting experiments were carried out at two different protein solution concentrations, 1

µM and 40 nM, respectively. The results were normalized, binned, and averaged for multiple tethers; see Figs. 2.2C and D. Again, in contrast to the linear curvature sorting observed for epsin ENTH, the measurements in Figs. 2.2C and 2.2D show significant deviations from linearity. Fig. 2.2C shows that for low values of the square root of

0 tension, the ratiometric parameter Ir / I r increases almost linearly for the case of 1 µM

protein solution concentration. As curvature is further increased, the sorting ratio

becomes nearly constant (Fig. 2.2C). At low protein solution concentration and low membrane tension (Fig. 2.2D), curvature sorting is significantly weaker compared to higher tensions (at the same solution concentration). For this concentration, while the membrane tension increases, the curvature/composition coupling also increases as

indicated by the larger slope of the fluorescence intensity ratios. We note that Figs. 2.2C

and D display relatively large standard deviations comparing different vesicles. The

27

sources for this variability may include differences in individual vesicle lipid

compositions. However, the main features of our measurements, i.e., non-linear sorting

and saturation of sorting at high membrane curvature and protein solution concentration,

were reproducible for individual vesicles. From fluorescence intensity values of the lipid

dye, measured on vesicle and tether, it is possible to estimate the radius of the tether ((18),

also see Materials and Methods). The results for our two solution conditions are shown in

Figs. 2.2E and F, for the same vesicles shown in Figs. 2.2C and D. The comparison of the experimental radii to those calculated assuming a bending stiffness of 0.8 ⋅ 10-19J (15) and absence of spontaneous curvature (dashed lines in Figs. 2.2E and F), reveals curvature generation at the higher, but not at the smaller protein solution concentration.

These curvature generation measurements, along with the curvature sorting results, were fitted with a theory (solid lines in Figs. 2.2C to F) detailed below.

3. ENBAR mobility on the tubular membrane is curvature dependent

In addition to the equilibrium curvature sorting measurements described above, we assessed curvature-dependent diffusion of ENBAR on tubular membranes via FRAP measurements. Figs. 2.3A and B show examples of fluorescence recovery after photobleaching of ENBAR on the tether membrane, demonstrating the mobility of

ENBAR on membranes. Individual measurements at a protein solution concentration of 1

µM were recorded for varying membrane tensions and analyzed as described in the materials and methods section. A one-dimensional diffusion model was then fit to the time-dependent recovery ratios. Figs. 2.3A and B show experimental results compared to fitted curves for smallest and largest membrane tensions, respectively. The diffusion 28

coefficient for the measurement displayed in Fig. 2.3A is 1.47 μm2/s, and the result for the data in Fig. 2.3B is 0.13 μm2/s. Quantitative photobleaching recovery measurements were obtained from image analysis of time lapse recordings of tether membrane fluorescence; see Fig. 2.3C. The results of FRAP measurements for a series of different membrane tensions are summarized in Fig. 2.3D. As membrane tension increases, the diffusion coefficient of ENBAR on membrane tethers decreases. As further discussed below, we hypothesize that the decrease in diffusion coefficients results from an increase in molecular crowding as the density of protein increases with rising curvature. We confirmed the hypothesis of protein density affecting diffusion by the following experiment.

The lateral mobility of ENBAR on tubular membranes was monitored by an alternative method that consisted of step-wise tether elongations. Membrane tethers previously equilibrated in the presence of ENBAR (1 μM) were rapidly (10 µm/s) extended by 10

μm, which resulted in a tether region with low protein coverage being pulled from the aspirated vesicle. ENBAR was observed to diffuse from the vesicle onto the tether (Fig.

2.4A) consistent with the photobleaching results discussed above; see Fig. 2.3C. This phenomenon could be reproduced several times by repeating the elongation process described above.

Furthermore, it is observed that for comparable membrane tensions, the diffusion of

ENBAR onto the tubular membrane after tether elongation (Fig. 2.4B) is significantly faster compared to diffusion observed after photobleaching (Fig. 2.4C). This observation supports our hypothesis that the lateral mobility of membrane-bound ENBAR depends on

29

the free area available for diffusion. In this view, diffusion kinetics at high lateral

tensions are slowed (Fig. 2.3 D) due to molecular crowding.

D. Discussion.

In the following sections we outline the derivation of a curvature sorting model that captures several of our experimental observations. We compare this model to our data, and discuss the possibility of curvature-induced phase transitions predicted by this model.

1. Introduction of a non-linear curvature/composition coupling model Classical analytical curvature/composition coupling models assume a linear coupling between local composition and local MC(19). Similar models have recently been used to interpret the partitioning of peripheral proteins in curvature gradients(15, 17-18). Our findings for the curvature partitioning of ENBAR shown in Figs. 2.2 C and D clearly deviate from linear coupling (note that in linear curvature/composition coupling models the curvature is proportional to Σ (17)). Thermodynamic terms in linear curvature/composition coupling models can be interpreted as terms resulting from second order Taylor

expansion in composition and curvature of the free energy(15, 17). In such models, the

coefficients of these expansions are evaluated for the thermodynamic reservoir (i.e., the

GUV) that pulled tethers are in contact with. In the following we replace the expansion

term squared in composition change by Γ, which is a function of fractional protein

coverage θ (ranging from zero to one), to define the tube free energy Ft:

2 κ  1   Ft = 2πRL  −θC p  + Σ + Γ(θ ) − fL  2  R   (9)

30

where R and L are tether radius and length, respectively, κ is membrane bending stiffness,

Cp is a spontaneous curvature of the membrane induced by protein binding, Σ is the lateral tension, and f is the pulling force acting on the tether. We note that this highly simplifying model neglects aspects such as the area difference elasticity(20), osmotic effects(21), membrane undulations, and the possibility of more than one protein binding mode(10). We also assume that the phenomenological spontaneous curvature Cp does not depend on membrane curvature.

In Eq. 9, the function Γ results from Legendre transform of a van der Waals free energy density f0 that describes the thermodynamics of the protein on the membrane:

µvesθ Γ(θ ) = f 0 (θ )− + Π ves (10) b where f0 is the mixing free energy density of a two-dimensional van der Waals gas:

k Tθ 1−θ  k Tθ θ 2 = − B − B − f0 ln  a 2 b  θ  b b . (11)

Here, kB is Boltzmann’s constant, T is temperature, b is the excluded area for protein coverage and a is a van der Waals interaction term (that here characterizes protein/protein interactions).

The function Γ shows a non-parabolic dependence on composition (as opposed to the usual Taylor expansion term(17-19)). This expansion term, which replaces Γ in Eq. 9, is written for the van der Waals model as follows:

1 2 kBT 2a 2 χ∆θ (12a), with χ = − + κC p , (12b) 2 θ (1−θ )2 b b2

31

where χ is the inverse osmotic compressibility of the van der Waals model. In Fig. 2.5 A,

Eq. 12a is compared to Eq. 10. We have for chemical potential and pressure of the van der Waals gas on a flat membrane:

  θ  θ  2aθ µ = k Tln ves  + ves  − ves + κC 2θ b ves B  1−θ  1−θ  b p ves   ves  ves  (13a)

k T θ θ 2 1 Π = B ves − a ves + κC 2θ 2 ves b 1−θ b2 2 p ves ves (13b)

μves is the (fixed) chemical potential of proteins bound to the vesicle (and in the aqueous

solution), and Πves a two-dimensional van der Waals pressure of the protein on the vesicle

(where curvature is assumed to be negligible). With this definition of Π, Σ is the lateral

tension in the vesicle membrane measured by micropipette aspiration.

Mechanical balance is obtained from minimization of Eq. 9 with respect to R:

κ 1 2 2 Σ = 2 − κC pθ − Γ(θ ) . (14) 2R 2 We note that in the absence of a reservoir (in which case Γ disappears), the familiar mechanical balance of a tube with spontaneous curvature is recovered from Eq. 14 (22-23).

Furthermore, in the absence of spontaneous curvature, the last two terms in Eq. 14 (and

therefore temperature and composition dependence) disappear at equilibrium, as required.

Using the chemical equilibrium condition obtained from Eq. 9 through minimization with

respect to θ, we can express Σ as a function of θ:

2 2 1  µves −κC pθb − (∂f0 ∂θ )b  1 Σ(θ ) =   − κC 2θ 2 − Γ(θ ) (15) κ   p 2  C pb  2 .

32

Eq. 15 is an analytically tractable relationship that can be compared to our experimental

data. Fig. 2.5 B compares Taylor expansion solution (leading to linear curvature sorting)

to the exact solution of the van der Waals model (which displays sigmoidal curvature

sorting), for identical parameters.

2. Comparison of analytical model to experimental data In the following we

demonstrate that our van der Waals curvature sorting model captures our experimental

observations, besides its apparent simplifications.

As mentioned above, the relative sorting ratio displayed in Figs. 2.2 C and D is a ratio of

the fluorescence intensities of proteins (Igreen) and of lipid probes (Ired) in the highly

0 0 curved tether (Ir= Igreen / Ired) normalized by the ratio I r ( I r = Ives-green / Ives-red) found on the vesicle. This normalized sorting ratio is equivalent to the ratio of coverage fractions on tether and vesicle(17), and can also be interpreted as the relative increase of protein density compared to the vesicle reservoir. A comparison of Figs. 2.2 C and D suggests

that the relative enrichment of ENBAR on the tubular membrane is smaller at larger solution concentration (1 μM, Fig. 2.2 C) compared to lower concentration (40 nM, Fig.

2.2 D). Fig. 2.5 C theoretically confirms this observation: for otherwise identical parameters of the model the relative enrichment increases with decreasing vesicle protein coverage fraction (which is related to the protein solution concentration by a binding isotherm; see (3)).

The van der Waals curvature sorting model contains four fit parameters. These are the interaction term a, the excluded area b, the spontaneous curvature Cp, as well as the vesicle coverage fraction θves that is related to μves; see Eq. 13a. The fit lines shown in Fig 33

2 result from simultaneously fitting of sorting and radius values (via Eqs. 14 and 15; see

Figs. 2.2 C and E as well as 2.2 D and F). We obtain spontaneous curvatures of 0.14 ±

0.007 nm-1 and 0.019 ± 0.0002 nm-1 for high and low protein concentration, respectively.

The fit values for b are in good agreement with the protein cross-section area(11); see the

supplementary material for all fit parameters and uncertainties.

3. Possibility of curvature-induced phase transitions Evidence from electron

microscopy imaging suggests that membrane tether regions covered by BAR domain

proteins can show differing, potentially coexisting, degrees of curvature(10, 24). We show

in the following that, as expected for any van der Waals type mixing model, our

curvature/composition coupling model predicts the existence of a first order phase

transition.

Equating to zero the determinant of the stability matrix resulting from Eq. 9 yields the

spinodal line (i.e., the local stability limit) for our model (see Fig. 2.6 A, thin black line).

a k T = B (16) b 2θ (1−θ )2

In Fig. 2.6 A, the spinodal is expressed as a function of lateral tension by solving Eq. 16

for θ (for a set of variable temperatures) and evaluating the associated lateral tensions by

means of Eq. 15. It is observed (and can be shown analytically) that the limit of stability is

reached when the slope of the curvature adsorption isotherm of Fig. 2.6 A is infinite, i.e.,

where ∂Σ1/ 2 ∂θ = 0 ; see solid dots in Fig. 2.6 A. The critical point of phase coexistence

(square on the tip of the spinodal line, Fig. 2.6 A), is found from Eq. 16 evaluated at the

critical composition θcrit = 1/3; acrit b = 27kBT 8 . 34

The phase boundary (i.e., the binodal line; see thick black line in Fig. 2.6) is obtained by

numerically solving the equations for chemical potentials of the tube, µt, and pulling force,

f:

µ = µ θ Σ µ = µ θ Σ θ Σ = θ Σ ves t ( 1, ) (17a) ves t ( 2 , ) (17b) f ( 1, ) f ( 2 , ) (17c)

for the three unknowns, θ1 and θ2 (the compositions of coexisting phases) and associated

lateral tension Σ, for given values of a, b, Cp, and μves. The first two conditions result from chemical equilibrium in the coexistence regime, and the last condition ensures mechanical balance (i.e., the pulling force on the membrane tether is equivalent in coexisting phases).

The pulling force is obtained from Eq. 9 as:

f 2π = 2κ Σ + Γ + 1 κC 2θ 2 − κC θ (18) ( 2 p ) p

Fig. 2.6 B demonstrates that a curvature-induced phase transition leads to a discontinuous

jump in tether radius (open dots in Fig. 2.6 B) associated with a discontinuous jump in

protein density (open dots on binodal line, Fig. 2.6 A). Furthermore, tiny changes in

curvature (from 17.6 to 18.1 nm) can lead to substantial protein density changes (from θ =

0.21 to 0.48, see Fig. 2.6). Our curvature sorting data do not yet reveal such a transition;

potentially because curvature changes during this transition may be hard to experimentally

resolve. It is, however, tempting to speculate that such curvature-induced phase transitions

might play the role of a curvature-dependent protein-density switch in processes that involve membrane deformation, such as the generation of tubular or vesicular membrane trafficking vehicles during endocytosis.

35

Unit conversion: For calculations based on the van der Waals curvature sorting model

introduced here, the following non-dimensionalized parameters (indicated by a hat) were used:

Σ R b κ a κ Σˆ = ⋅ R 2 ; Cˆ = C R ; Rˆ = ; bˆ = ⋅ ; aˆ = ⋅ (19) κ 0 p p 0 R k T 2 2 2 0 B R0 (kBT ) R0

Here, R0 is a reference length, and κ is the bending stiffness of the tether membrane.

36

Figures

Figure 2.1 Rat endophilin A1 NBAR domains (ENBAR) partition in curvature gradients generated by tether membranes pulled from GUVs.

A. Confocal xy images demonstrating Alexa Fluor 488 labeled ENBAR (green) enrichment on a tether pulled from micropipette-aspirated GUV membrane (red) with

composition 74% DOPC, 25% DOPG, 0.3% Texas Red-DHPE TR-DHPE and 0.7%

DSPE-Bio-PEG2000. Σ = 0.166 mN/m. 150 nM ENBAR in 33 mM NaCl, HEPES buffer

pH 7.4. Scale bar: 3 μm. Panels from up to down are images from protein channel, lipid

channel and merged channel, respectively.

B. Cross section confocal xz line-scan images of a membrane tether under varying

tensions demonstrating the curvature preference of ENBAR. Scale bar: 2 μm.

C. Demonstration of reversibility and equilibration for quantitative fluorescence

measurements of ENBAR curvature partitioning. Protein (green) and lipid probe (red)

37

fluorescence intensities measured during cyclic rapid membrane tension changes

[corresponding tension values in panel D.] Left axis, red squares, Ired; right axis, green diamonds, Igreen.

D. Ir = Igreen/Ired values for the data in panel C. (left axis, black diamonds), with indicated tension levels (right axis, blue squares).

38

Figure 2.2 ENBAR localization depends on membrane curvature.

A. Plot of separate green and red channel fluorescence intensities from images equivalent

to those shown in Fig. 2.1B for multiple tension levels from one tether membrane. While

the red (lipid) fluorescence intensity decreases as membrane tension is increased, the

green (protein) fluorescence increases. ENBAR concentration is 40 nM.

B. Plot of ratio of green and red channel fluorescence intensities for the data shown in Fig.

2 A.

C. Data from seven vesicles were binned (gray and black vertical error bars represent

0 standard deviations and standard errors of mean of I r , respectively; horizontal error bars show standard errors of mean of square root of tension). ENBAR concentration is 1 µM. 39

The gray solid line shows the fit with the van der Waals curvature sorting model

described in the main text.

D. Curvature sorting results as in panel C for ENBAR concentration of 40 nM. Seven vesicles were analyzed. The gray line shows a fit with the van der Waals sorting model.

E. & F. Tether radii changing as a function of membrane tension for ENBAR concentration of 1 µM (E) and 40 nM (F), respectively. The gray solid lines show fit results via the van der Waals model. The black dashed lines are the expected tube radii in the absence of protein.

40

Figure 2.3 ENBAR diffusion on membrane tethers measured via fluorescence

recovery after photobleaching (FRAP) slows with increasing curvature.

Bleached lengths amounted to 10 μm, and ENBAR concentration was 1 μM in 10 mM

NaCl. Diffusion coefficients were obtained from a one-dimensional diffusion model, and

continuous lines represent the fitting results.

A. Relative fluorescence recovery (defined in the methods section) of a membrane tether

for membrane tension Σ = 0.010 mN/m, D = 1.47 μm2/s.

B. Relative fluorescence recovery of a membrane tether for membrane tension Σ = 0.153

mN/m, D = 0.13 μm2/s.

C. Fluorescence images showing photobleaching recovery of a tether at membrane

tension Σ = 0.120 mN/m, D = 0.15 μm2/s. Scale bar: 5 μm.

D. Summary of ENBAR diffusion coefficients on tether membranes with varying degrees

of lateral tension and therefore varying degrees of curvature. Errors were determined as

explained in the materials and methods part.

41

Figure 2.4 Diffusion of ENBAR on membrane tethers is faster in tether elongation

experiments compared to FRAP experiments at similar membrane tension.

Elongation as well as bleached lengths are 10 μm and ENBAR concentration is 1 μM in

solution of 10 mM NaCl. Diffusion coefficients were obtained from a one-dimensional diffusion model, and continuous lines represent the fit results.

A. Fluorescence images showing time dependence of protein fluorescence increase of a tether membrane stretch that was rapidly pulled from a GUV with a tether that had previously been equilibrated with ENBAR at membrane tension Σ = 0.110 mN/m; D =

1.86 μm2/s. Scale bar: 3 μm.

B. Fluorescence recovery after tether elongation. Fit results in diffusion coefficient of D

= 2.17 μm2/s at membrane tension Σ = 0.076 mN/m.

C. Fluorescence recovery after photobleaching. Fit results in diffusion coefficient of D =

0.28 μm2/s at membrane tension Σ = 0.068 mN/m.

42

Figure 2.5 Comparison of curvature/composition coupling model to experimental

data.

2 2 A. Black solid line: function (Γ + κC pθ / 2) using a vesicle coverage of θves = 0.093,

ˆ ˆ and non-dimensionalized parameters aˆ = 0, b = 0.00096, and C p = 167 (equivalent to b

2 -1 = 50 nm and Cp = 1/6 nm at room temperature), which are values corresponding to a

stable (i.e., homogenous) regime at the equilibrium concentration. As expected, the

harmonic approximation (gray dashed line) is accurate near the vesicle (equilibrium)

coverage.

B. Plot of the ratio θtether/θves as a function of lateral tension. The exact expression is shown as black solid line and the gray dashed line shows the linear sorting resulting from the Taylor expansion approximation for the same set of parameters as in panel A.

ˆ ˆ C. Van der Waals model isotherms evaluated for the same values for aˆ , b , and C p as in panel A but different vesicle coverage fractions of 0.03 (light gray), 0.05 (medium gray) and 0.093 (black).

43

Figure 2.6 Van der Waals curvature/composition coupling model predicts

curvature-driven phase transition.

A. Binodal (thick black line) and spinodal (thin black line) for the van der Waals

curvature/composition coupling model, for varying molecular interaction parameter aˆ

ˆ ˆ (equivalent to varying temperature, see Eq. 16). Parameters bcrit , θves and C p were

0.0012, 0.0189 and 19.513, respectively, close to the fitting parameters for experimental

outcomes of curvature sorting and radius measurements at protein concentration at 40 nM.

Open square refers to the critical point where aˆ = aˆcrit . The intersections of the curvature

isotherm (light gray line) and the binodal curve for aˆ > aˆcrit are the phase coexistence

points (open circles). The dashed line represents a branch on the isotherm that lies within

the coexistence regime. The intersections of the isotherm and the spinodal line represent

the stability limits (solid circles). Isotherms with 0 < aˆ < aˆcrit (darker gray lines) are stable over the entire curvature range. 44

B. Relationship between inverse tether radius (curvature) and membrane tension. The branch between the solid black points corresponds to the instable region shown in panel A.

The open circles are coexistence points (corresponding to open dots on the binodal line in panel A). At the phase boundary, the tether radius shows a discontinuous jump.

45

[ENBAR] = 1 μM (Figs. 2.2 C and E) Non-dimensional dimensional 2 a 0.00024±0.0003 12.34±16.5 kBT∙nm b 0.0011±0.00009 56.54±4.4 nm2 θves 0.096±0.004 0.096±0.004 -1 Cp 139.94±7.2 0.14±0.007 nm [ENBAR] = 40 nM (Figs. 2.2 D and F) Non-dimensional dimensional 2 a 0.0045±0.0001 231.32±7.1 kBT∙nm b 0.0013±0.00005 66.83±2.8 nm2 θves 0.019±0.0005 0.019±0.0005 -1 Cp 18.94±0.2 0.019±0.0002 nm

Table 1. Fit parameters for protein sorting and tether radius measurements.

The table lists the fit parameters resulting from simultaneously comparing the van der

Waals curvature sorting model to the experimental protein sorting (Figs. 2.2 C and D) and curvature generation (Figs. 2.2 E and F) data. The uncertainties are estimated as described above.

46

E. Reference.

1. Hokanson, D. E., and E. M. Ostap. 2006. Myo1c binds tightly and specifically to phosphatidylinositol 4,5-bisphosphate and inositol 1,4,5-trisphosphate. P Natl Acad Sci USA 103:3118-3123. 2. Mathivet, L., S. Cribier, and P. F. Devaux. 1996. Shape Change and Physical Properties of Giant Phospholipid Vesicles Prepared in the Presence of an AC Electric Field. Biophysical Journal 70:1112 - 1121. 3. Zhu, C., S. L. Das, and T. Baumgart. 2012. Nonlinear sorting, curvature generation, and crowding of endophilin NBAR on tubular membranes. Biophysical Journal 102:1837-1845. 4. Crank J. 2004. The Mathematics of Diffusion. Oxford University Press, New York. 5. Sorre, B., A. Callan-Jones, J. Manzi, B. Goud, J. Prost, P. Bassereau, and A. Roux. 2012. Nature of curvature coupling of amphiphysin with membranes depends on its bound density. Proceedings of the National Academy of Science of the United States of America 109:173-178. 6. Bevington, P. R., and D. K. Robinson. 2003. Data reduction and error analysis for the physical sciences. McGraw-Hill, Boston. 7. Tian, A., B. R. Capraro, C. Esposito, and T. Baumgart. 2009. Bending stiffness depends on curvature of ternary lipid mixture tubular membranes. Biophysical Journal 97:1636-1646. 8. Tian, A. W., B. R. Capraro, C. Esposito, and T. Baumgart. 2009. Bending Stiffness Depends on Curvature of Ternary Lipid Mixture Tubular Membranes. Biophysical Journal 97:1636-1646. 9. Masuda, M., S. Takeda, M. Sone, T. Ohki, H. Mori, Y. Kamioka, and N. Mochizuki. 2006. Endophilin BAR domain drives membrane curvature by two newly identified structure-based mechanisms. Embo Journal 25:2889-2897. 10. Mizuno, N., C. C. Jao, R. Langen, and A. C. Steven. 2010. Multiple modes of endophilin-mediated conversion of lipid vesicles into coated tubes: implications for synaptic endocytosis. Journal of Biological Chemistry 285:23351-23358. 11. Gallop, J. L., C. C. Jao, H. M. Kent, P. J. G. Butler, P. R. Evans, R. Langen, and H. T McMahon. 2006. Mechanism of endophilin NBAR domain-mediated membrane curvature. Embo Journal 25:2898-2910. 12. Cui, H. S., G. S. Ayton, and G. A. Voth. 2009. Membrane Binding by the Endophilin NBAR Domain. Biophysical Journal 97:2746-2753. 13. Arkhipov, A., Y. Yin, and K. Schulten. 2008. Four-scale description of membrane sculpting by BAR domains. Biophys J 95:2806-2821. 14. Blood, P. D., and G. A. Voth. 2006. Direct observation of Bin/amphiphysin/Rvs (BAR) domain-induced membrane curvature by means of molecular dynamics simulations. P Natl Acad Sci USA 103:15068-15072. 15. Tian, A., and T. Baumgart. 2009. Sorting of lipids and proteins in membrane curvature gradients. Biophysical Journal 96:2676 - 2688.

47

16. Baumgart, T., B. R. Capraro, C. Zhu, and S. L. Das. 2011. Thermodynamics and Mechanics of Membrane Curvature Generation and Sensing by Proteins and Lipids. Annual Review of Physical Chemistry, Vol 62 62:483-506. 17. Capraro, B. R., Y. Yoon, W. Cho, and T. Baumgart. 2010. Curvature sensing by the epsin N-terminal homology (ENTH) domain measured on cylindrical lipid membrane tethers. Journal of the American Chemical Society 132:1200 - 1201. 18. Sorre, B., A. Callan-Jones, J. Manzi, B. Goud, J. Prost, P. Bassereau, and A. Roux. 2012. Nature of curvature coupling of amphiphysin with membranes depends on its bound density. P Natl Acad Sci USA 109:173-178. 19. Leibler S. 1986. Curvature Instability in Membranes. Journal de Physique 47:507 - 516. 20. Seifert U., Berndl K., and Lipowsky R. 1991. Shape Transformation of Vesicles: Phase Diagram for Spontaneous-curvature and bilayer-coupling models. Physical Review A 44:1182 - 1202. 21. Lipowsky, R., and H. G. Dobereiner. 1998. Vesicles in contact with nanoparticles and colloids. Europhysics Letters 43:219-225. 22. Bukman, D. J., J. H. Yao, and M. Wortis. 1996. Stability of cylindrical vesicles under axial tension. Physical Review E 54:5463-5468. 23. Li, Y. H., R. Lipowsky, and R. Dimova. 2011. Membrane nanotubes induced by aqueous phase separation and stabilized by spontaneous curvature. P Natl Acad Sci USA 108:4731-4736. 24. Henne, W. M., H. M. Kent, M. G. J. Ford, B. G. Hegde, O. Daumke, P. J. G. Butler, R. Mittal, R. Langen, P. R. Evans, and H. T. McMahon. 2007. Structure and analysis of FCHo2F-BAR domain: A dimerizing and membrane recruitment module that effects membrane curvature. Structure 15:839-852.

48

Chapter 3: Membrane interacting helices influencing curvature sorting and

membrane kinetics of endophilin NBAR

A. Background

Membrane insertion amphipathic helices are thought to play an important role in NBAR

domain proteins membrane interactions(1-2). For endophilin, both N-terminal and insert helices (H0 and H1I) are believed to insert into the membrane to facilitate membrane bending (3-10). Furthermore, H0 has been suggested to mediate the oligormerization of endophilin on the membrane (2, 11). However, there are historically controversial opinions about the importance of insertion helices on NBAR domain protein membrane curvature (MC) sensing and generation: For the curvature sorting aspect, single liposome curvature assay suggested H0 insertion was necessary for NBAR domain MC sensing(12), while Peter et al(13) believed BAR domain alone could sense MC; For the curvature generation ability, Simulations(14) and EM(8) results suggested scaffolding, rather H0 helix insertion, promoted NBAR domains membrane bending. But other researchers stated H0 could bend membrane solely, and accelerated BAR domain membrane interaction(6-7). These contradictory opinions call for more investigations.

In this chapter, we investigated the influences of membrane insertion amphipathic helices on NBAR domain membrane curvature sensing abilities, by comparing the helices deletion mutants with the WT endophilin NBAR domain (ENBAR). We found that both

N-terminal helix and insert helix have indispensible contributions to ENBAR curvature sorting. Besides these equilibrium studies, we explored the effects of membrane insertion amphipathic helices on ENBAR membrane dissociation kinetics. We observed for both

WT and the mutant, their membrane dissociation time were positively correlated with 49

their membrane densities, implying the formation of oligomers on the membrane. We

also found irreversible binding fractions during the kinetics measurement. These

remaining components could be the linear aggregation(15) of ENBAR on the membrane.

B. Materials and methods

1. Materials:

Proteins: recombinant rat endophilin A1 NBAR plasmid was generously provided by Pr.

De Camilli, Yale University, USA. Recombinant rat endophilin A1 NBAR ΔH1I plasmid

was generously provided by Pr. Mochizuki, National Cerebral and Cardiovascular Center

Research Institute, Japan. Recombinant rat endophilin A1 NBAR ΔH0 plasmid was generated by our lab member Dr. Hsieh, through site-direct mutagenesis. The expression and purification of recombinant proteins are described in Appendix A. The proteins are labeled with Alexa Fluor-488 as described in Appendix A.

Lipid: All synthetic lipids were purchased from Avanti. The lipid composition used for the sorting experiments and single giant unilamellar vesicle (GUV) transfer experiments was

DOPC 64%, DOPG 25%, PI(4,5)P2 10%, DSPE-Bio-PEG2000 0.7% and TR-DPHE

0.3%. GUV formation is described in Chapter 2B.

2. Methods:

Curvature sorting experiments and reversibility measurements are described in Chapter

2B.

50

Single GUV transfer experiments:

Single GUV transfer experiments were performed as previously described(16). Briefly,

the preparation of measurement chambers and aspiration micropipettes were similar to

those in the sorting experiment. The outer micropipettes which used for transfer were

cutted from capillaries (World Precision Instruments Inc., Sarasota, FL). The proteins and

GUVs were incubated in Eppendorfs for around 30 min at room temperature. Then

around 150 µl of incubated sample was put in the measurement chamber. An aspiration

pipette covered with an outer capillary was inserted into the measurement chamber to

aspirate a protein-covered GUV. Then the outer capillary was adjusted to cover the vesicle with solution around. Subsequently the vesicle transfer was carried out by the micro manipulator and moved to the dilution buffer which was free of protein. After the transfer the outer capillary was removed to expose the vesicle in the buffer. The protein dissociation process on the GUV was recorded by a confocal microscope (17).

C. Results

1. Membrane insertion amphipathic helices are essential for ENBAR curvature

sensing ability.

The physicochemical functions of amphipathic helices in vitro were investigated by comparing ENBAR WT with its two mutants: ENBAR ΔH0 and ENBAR ΔH1I. I first examine how the curvature sensing ability of ENBAR is affected by its membrane insertion helices. Fig 3.1A shows qualitative curvature partitioning preferences of

ENBAR WT (upper row), ENBAR ΔH1I (middle row) and ENBAR ΔH0 (lower row).

For all three variants, their fluorescence signals on the tether (green) are almost the same 51

as those on the GUVs, while the signals from lipid on the tether are significantly weaker than those on the GUVs. These observations demonstrate curvature sorting of all three variants, because the relative enrichment of the protein molecules on the high curvature membrane compared to the lipid molecules.

To gain an in-depth understanding of the contributions of amphipathic helices in ENBAR curvature sensing, quantitative curvature sorting measurements were performed. In order to be able to interpret our results, we have to guarantee that all of our characterizations are done under the equilibrium circumstances. To accomplish this, the reversibility of the partitioning of all three proteins on membrane tethers were examined, see (fig 3.2). Both

ENBAR WT and the two mutants are able to vary their degree of curvature partitioning according to the degree of the membrane tension.

Having demonstrated reversibility, quantitative curvature sorting measurements of

ENBAR WT and mutants were then performed at identical protein solution concentrations 40nM (fig 3.1B). Clearly, the WT curvature sorting ability is stronger compared to both mutants, while the sorting degrees of both mutants are similar (no statistically significant difference).

Considering the nature of curvature coupling of NBAR protein with membrane depending on its membrane density (17-19), it is of interest to compare the curvature sorting abilities of ENBAR WT and mutants on the same protein membrane density level.

Therefore, to obtain the density of proteins on the membrane, GUV binding isotherm experiments were determined for all three proteins (fig 3.3A), which help us figure out appropriate protein solution concentrations that result in similar membrane binding densities for proteins with different membrane binding. Curvature sorting quantifications 52

were then carried out similar with above, but with the same protein membrane density instead of the same protein solution concentration. The results are shown in fig 3.3B. The

WT protein still shows stronger curvature sorting abilities than both mutants. Again, the sorting abilities for the mutants are found to be indistinguishable (no statistically significant difference). These results also suggest that deletion of any membrane insertion amphipathic helices would decrease the sorting efficiency of ENBAR.

2. Results from membrane dissociation kinetics of ENBAR WT and ΔH1I implied

higher order oligmerization on the membrane.

To further understand the influence of membrane insertion amphipathic helices on

ENBAR membrane dissociation kinetics, we investigated kinetic aspects of membrane binding. The membrane dissociation kinetics of ENBAR WT and ENBAR ΔH1I were measured via single GUV transfer experiment (16). Fig 3.4A shows an example of a fluorescent image of an aspirated GUV chosen to be transferred. The fluorescence intensity on the GUV is fitted by a Gaussian ring as shown in fig 3.4B to extract the intensity value, which later is converted to the protein density on the membrane and used in the following analysis. Typically the ENBAR � H1I shows a single exponential decay

(fig 3.4C) when it dissociates from the membrane.

Our group has previous found a positive correlation between protein membrane dissociation time and protein density on membrane, suggesting that proteins oligomerize on the membrane(16). We next asked whether the amphipathic helix is mediating endophilin membrane oligomerization. For this purpose, the ENBAR WT and ENBAR

ΔH1I membrane dissociation time were investigated as a function of protein density. Fig 53

3.5A shows that ENBAR WT has stronger oligomerization tendency than ENBAR ΔH1I,

since the slope of linear regression of the WT protein is significantly higher than that of the mutant (P<0.05*), which implies that H1I helix might also contribute to membrane oliogmerization, besides H0 helix (16).

Another interesting observation is the existence of what appear to be irreversible binding

fractions for both ENBAR WT and ENBAR ΔH1I during their dissociation process, as

shown by the observation of a clear plateau in the fluorescence trace in fig 3.4C.

Furthermore, the amount of this irreversible binding fraction is proportional to the total

protein density on the membrane (fig 3.5B). This means the percentage of the remaining

component/ irreversible binding fraction is not a function of protein membrane density;

instead, it is an independent parameter For ENBAR WT and ΔH1I (fig 3.5C), it is 32.5 ±

9.1% and 42.6 ± 6.2%, respectively (uncertainty is STD). In order to test the possibility that the irreversible binding fractions are due to membrane nano-tubulation below the optical resolution, we carried out EM experiment with the same protein to lipid ratio

(with our lab member Tingting Wu’s help) and the same all other experimental

conditions as in the single GUV transfer experiment. We find that there are no membrane

tubulation at this condition (fig 3.5D), which exclude that possibility.

D. Discussion

Membrane insertion amphipathic helices generally exist in many cellular transportation

related proteins, besides BAR domain containing proteins(20-21). The function of amphipathic helices however, is historically controversial (as discussed in background

part). Further, a recent study proposed that amphipathic helices are a driving force for 54

membrane scission while the BAR domain scaffolds inhibit fission(1). In all, it is clear that amphipathic helices play an essential role in CME; while the exact mechanism of the

function of these helices is not very well understood yet.

Here we contributed an improved understanding of the functions of amphipathic helices,

via monitoring ENBAR WT and its helices deletion mutants curvature sensing,

membrane dissociation and possible oligomerziation abilities. The curvature sorting

results we obtained indicate: a) helix insertion into membranes is indispensable to

ENBAR curvature sensing indispensable. This conclusion is consistent with that fact that

both mutants possess weaker curvature sorting abilities than WT; b) contributions from

different helices to curvature sensing are experimentally indistinguishable by the methods

used here. The insertion depths of these two helices are different(4). Besides, EM

tubulation assays suggest the helices deletion mutants curvature generation abilities are

different (4, 22-23). The possible reasons for the plausible discrepancies between that

they observed differences between these two helices and that we didn’t find significant

differences between these two helices could be: a) the differences for the H0 and H1I

helices are mainly on their abilities to initiating the curvature (the differences in their

tabulation abilities in EM experiments (22)), instead of sensing the curvature; b) the

discrepancies in protein/lipid ratio(23) or protein membrane density(19) in different

experiments could couple with the experimental outcomes in various assays; c) further

investigation is needed for the relationship between the physical parameters of the helix

insertion (e.x. insertion depth, angle, etc) and its curvature sensing/ initiation ability.

The kinetic aspect of the NBAR domain protein membrane interaction is another

intriguing realm in the CME process. Interestingly, a recent publication(16) suggests the 55

oligomerzation inclination of NBAR domain proteins could be resolved through the kinetic experiments. Our results states that both ENBAR WT and ΔH1I could form higher order structures on the membrane, while the former has stronger oligomerization tendency. That suggests that the H1I helix might also participate in the oligomerization process of ENBAR, while more investigation is need to fully understand the mechanisms involved. Besides, we found irreversible binding fractions for both ENBAR WT and

ΔH1I on the membrane. The amount of this remaining fraction is proportional to the total protein density on the membrane; therefore its percentage remains a constant in the protein membrane density gradient. We have excluded the possibility that this irreversible binding is due to nano-tubulars below the optical resolution. We hypothesize that this phenomenon implicates lattice structure/ higher order aggregation of proteins formed on the membrane(15).

56

Figures

Figure 3.1: Curvature preference partitioning of ENBAR WT, ENBAR ΔH1I and

ENBAR ΔH0.

A. Fluorescent images of partitioning of all three variants. Lipids are labeled with

TR-DHPE (red); Proteins are conjugated with Alexa Fluor-488 (green). The aspirated vesicle is on the right of the image, connecting a pulled horizontal highly-curved tube.

While fluorescence signals from lipid are weaker on the tethers compared to the vesicles

(due to microscope focal volume being larger than the tube radii), the signals from the protein are comparable among tube and vesicle. This implies curvature preferences of these proteins. Scale bar: 5 μm.

B. Comparison of curvature sensing abilities of ENBAR WT and mutants at the same protein solution concentration. Fluorescence intensity values are obtained from the cross section areas of the pulled tethers(17).White circle, black circle and white square are for

ENBAR WT, ENBAR ΔH0 and ENBAR ΔH1I, respectively. Data are binned from 5 vesicles for each protein. Gray and black error bars are STD and SEM, respectively.

57

Figure 3.2: Demonstration of ENBAR WT and mutants curvature sorting reversibilities. Panel A&B, C&D and E&F are for ENBAR WT, ENBAR ΔH0 and

ENBAR ΔH1I, respectively. Panel A, C and E are lipid (red circle) and protein (green square) channel fluorescence intensities from the cross section areas of tether subjected to membrane tension change cycle, for three variants, respectively; panel B, D and F are the corresponding membrane tension (blue square) and sorting ratio (black circle) changes.

Time intervals between tension changes were determined based on the time of real pressure change.

58

Figure 3.3: Comparison of curvature sensing abilities of ENBAR WT and mutants at the same protein membrane density.

White circle, black circle and white square are for ENBAR WT, ENBAR ΔH0 and

ENBAR ΔH1I, respectively.

A. GUV Isotherm absorption experiments for all three variants. The data points are fitted

with Langmuir isotherm absorption equation (black curves). For each data point, 5

vesicles are measured. Gray and black error bars are STD and SEM, respectively.

B. The quantification of curvature sorting of all three variants at the same protein

membrane density. Data are binned from 5 vesicles for each protein. Gray and black error bars are STD and SEM, respectively.

59

Figure 3.4: ENBAR ΔH1I dissociation follows single exponential decay.

A. A representative fluorescence image of an aspirated GUV chosen to be transferred.

Scale bar: 5 μm.

B. The Gaussian ring used to extract the fluorescence intensity value on the GUV.

C. A typical dissociation trace of ENBAR ΔH1I. The black dots are data points and the gray line is single exponential decay fitting which gives a dissociation time of 422.8 ±

20.5s (uncertainty is from fitting standard error).

60

Figure 3.5: Monitoring ENBAR WT and ΔH1I membrane dissociation processes at different membrane densities.

Black circle and white (blue) square are for ENBAR WT andΔH1I respectively. Error bars are standard errors for the fitting parameters. The black lines are linear regressions.

A. Protein membrane dissociation time is positively correlated with protein membrane density for both proteins.

B. The amount of remaining component is positively correlated with protein membrane density for both proteins.

C. Presenting the results in B. in a percentage manner.

D. The representative EM image shows no membrane tubulation induced by proteins in the conditions which the experiments in panel A carried out. Scale bar: 200nm.

61

E. References:

1. Boucrot, E., A. Pick, G. Camdere, N. Liska, E. Evergren, H. T. McMahon, and M. M. Kozlov. 2012. Membrane Fission Is Promoted by Insertion of Amphipathic Helices and Is Restricted by Crescent BAR Domains. Cell 149:124-136. 2. Mim, C., H. Cui, Joseph A. Gawronski-Salerno, A. Frost, E. Lyman, Gregory A. Voth, and Vinzenz M. Unger. 2012. Structural Basis of Membrane Bending by the N-BAR Protein Endophilin. Cell 149:137-145. 3. McMahon, H. T., and J. L. Gallop. 2005. Membrane curvature and mechanisms of dynamic cell membrane remodelling. Nature 438:590-596. 4. Gallop, J. L., C. C. Jao, H. M. Kent, P. J. G. Butler, P. R. Evans, R. Langen, and H. T. McMahon. 2006. Mechanism of endophilin N-BAR domain mediated membrane curvature. The EMBO Journal 25:2898-2910. 5. Arkhipov, A., Y. Yin, and K. Schulten. 2008. Four-scale description of membrane sculpting by BAR domains. Biophysical journal 95:2806-2821. 6. Blood, P. D., R. D. Swenson, and G. A. Voth. 2008. Factors influencing local membrane curvature induction by N-BAR domains as revealed by molecular dynamics simulations. Biophysical journal 95:1866-1876. 7. Campelo, F., H. T. McMahon, and M. M. Kozlov. 2008. The hydrophobic insertion mechanism of membrane curvature generation by proteins. Biophysical journal 95:2325-2339. 8. Fernandes, F., L. M. S. Loura, F. J. Chichon, J. L. Carrascosa, A. Fedorov, and M. Prieto. 2008. Role of helix 0 of the N-BAR domain in membrane curvature generation. Biophysical Journal 94:3065-3073. 9. Cui, H., G. S. Ayton, and G. A. Voth. 2009. Membrane binding by the endophilin N-BAR domain. Biophysical Journal 97:2746-2753. 10. Hatzakis, N. S., V. K. Bhatia, J. Larsen, K. L. Madsen, P. Bolinger, A. H. Kunding, J. Castillo, U. H. Gether, P., and D. Stamou. 2009. How curved membranes recruit amphipathic helices and protein anchoring motifs. Nature Cell Biology 5:835-841. 11. Gallop, J. L., and H. T. McMahon. 2005. BAR domains and membrane curvature: bringing your curves to the BAR. Biochem. Soc. Symp. 72:223-231. 12. Bhatia, V. K., K. L. Madsen, P. Bolinger, A. Kunding, P. Hedegard, U. Gether, and D. Stamou. 2009. Amphipathic motifs in BAR domains are essential for membrane curvature sensing. The EMBO Journal 28:3303-3314. 13. Peter, B. J., H. M. Kent, I. G. Mills, Y. Vallis, P. J. G. Butler, P. R. Evans, and H. T. McMahon. 2004. BAR domains as sensors of membrane curvature: the amphiphysin BAR structure. Science 303:495-499. 14. Arkhipov, A., Y. Yin, and K. Schulten. 2009. Membrane-bending mechanism of amphiphysin N-BAR domains. Biophysical journal 97:2727-2735. 15. Simunovic, M., A. Srivastava, and G. A. Voth. 2013. Linear aggregation of proteins on the membrane as a prelude to membrane remodeling. Proceedings of the National Academy of Science of the United States of America 110:20396-20401.

62

16. Capraro, B. R., Z. Shi, T. Wu, Z. Chen, J. M. Dunn, E. Rhoades, and T. Baumgart. 2013. Kinetics of endophilin N-BAR domain dimerization and membrane interactions. The Journal of Biological Chemistry 288:12533-12543. 17. Zhu, C., S. L. Das, and T. Baumgart. 2012. Nonlinear sorting, curvature generation, and crowding of endophilin N-BAR on tubular membranes. Biophysical Journal 102:1837-1845. 18. Heinrich, M., B. R. Capraro, A. Tian, J. M. Isas, R. Langen, and T. Baumgart. 2010. Quantifying membrane curvature generation of drosophila amphiphysin N-BAR domains. J. Phys. Chem. Lett. 1:3401. 19. Sorre, B., A. Callan-Jones, J. Manzi, B. Goud, J. Prost, P. Bassereau, and A. Roux. 2011. Nature of curvature coupling of amphiphysin with membranes depends on its bound density. Proceedings of the National Academy of Sciences 109:173-178. 20. Ford, M. G. J., I. G. Mills, B. J. Peter, Y. Vallis, G. J. K. Praefcke, P. R. Evans, and H. T. McMahon. 2002. Curvature of clathrin-coated pits driven by epsin. Nature 419:361-366. 21. Lee, M. C. S., L. Orci, S. Hamamoto, E. Futai, M. Ravazzola, and R. Schekman. 2005. Sar1p N-terminal helix initiates membrane curvature and completes the fission of a COPII vesicle. Cell 122:605-617. 22. Masuda, M., S. Takeda, M. Sone, T. Ohki, H. Mori, Y. Kamioka, and N. Mochizuki. 2006. Endophilin BAR domain drives membrane curvature by two newly identified structure-based mechanisms. The EMBO Journal 25:2889-2897. 23. Jao, C. C., B. G. Hegde, J. L. Gallop, P. B. Hegde, H. T. McMahon, I. S. Haworth, and R. Langen. 2010. Roles of Amphipathic Helices and the Bin/Amphiphysin/Rvs (BAR) Domain of Endophilin in Membrane Curvature Generation. The Journal of Biological Chemistry 285:20164-20170.

63

Chapter 4: Endophilin full length H0/ SH3 intradimer / intermolecular autoinhibition

A. Background

Endophilin is a peripherally binding membrane protein involved in clathrin-mediated endocytosis (CME) (1-3). Endophilin is consisted by N-terminal H0 helix, BAR domain, a linker region and a Src Homology3 (SH3) domain(4). The SH3 domain of endophilin regulates CME by controlling the recruitment of dynamin and synaptojanin(5-6), which are the downstream binding partners of endophilin and important players in CME.

However, besides these known intra/ intermolecular interaction of SH3 with the endophilin N-BAR domain itself is not very well understood. An SAXS study(7) implied that the endophilin SH3 domain would interact with its own BAR domain in solution. A similar hypothesis was proposed by a study that revealed an autoinhibition mechanism for syndapin, an FBAR domain containing protein(8). However, recent theoretical research suggested that endophilin was auto-inhibited through its H0/ SH3 interaction, which would assist the stabilization of helix structure of H0 through the presence of SH3 domain in solution(9). This discrepancy about either H0 or BAR domain interacting with

SH3 domain calls for more experimental results for studying the intra/ intermolecular mechanism of endophilin.

In this chapter we investigated the intradimer / intermolecular interaction of endophilin by monitoring its H0/SH3 interaction. Förster resonance energy transfer (FRET) experiments were carried out between H0 and SH3 domains of endophilin. Quenching of fluorescence emission spectra of both donor and acceptor were observed. We then aimed to use Trp quenching as an approach to study H0/SH3 interactions. Fluorescence spectra 64

between SH3 domain at varying concentrations and H0 WT/ H0 F10W at a constant

concentration were obtained. Both peptides induced Trp quenching which increases with

increasing SH3 concentration. All these three kinds of experimental results support the

hypothesis that endophilin is auto-inhibited in solution via an H0/SH3 domain interaction.

This provides the first experimental evidence to reveal the intramolecular mechanism of endophilin and shed light on studying its functions in CME.

B. Materials and methods

1. Materials:

Proteins: recombinant rat endophilin A1 SH3 plasmid (245-352aa) was generously

provided by Pr. Haucke, Leibniz Institut für Molekulare Pharmakologie & Freie

Universität Berlin, Germany. Expression and purification of protein were as described in

Appendix A.

Peptide: peptide H0 WT (1-22aa, rat endophilin A1) and H0 F10FCN were synthesized by

Culik, Department of Chemistry, University of Pennsylvania. Peptide H0 F10W (10) was

purchased from Biological Chemistry Resource Center, Department of Chemistry,

University of Pennsylvania.

2. Methods:

Fluorescence measurement: Peptides were dissolved in HN15T buffer (Hepes 20mM,

NaCl 150mM, TCEP 1mM, PH 7.4). Proteins were freshly thawed from aliquots stored at

-80°C, and ultracentrifuged before usage. The fluorescence cuvette and Eppendorf tubes

for incubation were passivated with casein solution (0.5% w/v) and rinsed with HN15T 65

buffer before usage. All fluorescence spectra were obtained from steady state

measurements. Samples were incubated in Eppendorf tubes for around 15mins at room temperature (23 ±2°C) before measurements. Fluorescence spectra were taken by a Cary

Eclipse fluorometer with a Peltier-controlled temperature block, at room temperature.

H0 F10FCN/SH3 FRET measurements: The excitation wavelengths for H0 F10FCN/SH3

FRET measurements were 240nm. The donor (H0 F10FCN) to acceptor (SH3) ratio was

15:1 or 2:1.

Fluorescence measurements: Fluorescence experiments were carried out using an

excitation wavelength of 280nm. The peptide (H0 F10W or H0 WT) concentrations were

kept at 400 μM, they remained constant during the titration with varying SH3 domain

concentrations.

C. Results

1. H0F10FCN/ SH3 FRET experiments showed both donor and acceptor emission

decreasing.

Previous simulation results(9) suggested that an H0/SH3 interaction induced

autoinhibition of endophilin, and that this was largely due to the hydrophobic interaction

between H0 and SH3 domain. Specially, the Phe10 in H0 would favor to bury itself in the

peptide binding groove of the SH3 domain (fig 4.1A). The simulation results also showed

the strongest hydrophobic interactions existing between Phe10 and Trp327/ Phe338(9).

66

Therefore we wondered whether we could find experimental evidence for the interaction

between separated H0 and SH3 in solution through this hydrophobic interface.

A PheCN-Trp FRET pair was used to monitor the H0/SH3 interaction, by substituting

Phe10 by PheCN10 (structure in figure 4.1B) in the H0 helix (H0 F10FCN). This FRET

pair has a Förster distance of 16 angstroms (11). The simulation study (9) suggested the

distance between Phe10 in H0 andTrp327 in SH3 domain was around 5 angstroms in the

bound state, which gave us feasibility of choosing this FRET pair. Steady state FRET

measurements were performed with a donor to acceptor ratio of 15:1. Results are shown in figure 4.1C, which shows the fluorescence emission spectra of H0 F10FCN only, SH3

only, and the spectrum of incubated H0 F10FCN with SH3. Interestingly, the results

display simultaneous decrease of donor emission and acceptor emission intensities, which

is different from an ordinary FRET response which shows decreasing donor and

increasing acceptor emission intensities. More direct comparison of results in fig 4.1C is

shown in fig 4.1D. The summary of H0 F10FCN and SH3 separated spectrum results in a

clear comparison with the spectrum obtained from incubated samples. The quenching of

the donor emission intensity suggests the presence of FRET.

Non-linearity of H0 F10FCN fluorescence intensity-concentration relationships was

observed. Therefore the concentration range of H0 F10FCN was determined, within which

its intensity was proportional to its concentration. The results are shown in fig. 4.2.

Based on the H0 F10FCN concentration range we determined, steady state FRET experiments were performed using H0 F10FCN and SH3 at different concentrations but

maintaining the same donor to acceptor ratio (fig. 4.3). Again, increasing quenching of

67

both donor and acceptor emission intensities with increasing concentrations was observed.

It has been reported that quenching of the intrinsic Trp in SH3 or SH2 domain occurring

with binding of ligands (12-13). Therefore, I was wondering that whether the decrease of

Trp signal was due to the H0 helix binding induced Trp quenching, which overcame the

sensitized emission caused by FRET. To test this hypothesis, I performed the following

fluorescence experiments.

2. Trp in SH3 quenching induced by H0 F10W or H0 WT.

Since I was planning to use the signal of the intrinsic Trp in SH3 as a probe for studying

the binding between H0 and SH3, I first determined the linear regime of the Trp in SH3

fluorescence intensity-concentration relationship to validate our research (fig. 4.4).

In order to amplify the magnitude of Trp quenching during H0/SH3 interaction for better

signal to noise ratio, I introduced a Trp at position 10 in the H0 helix (through the

mutation F10W). A fluorescence titration between H0 F10W and SH3 was carried out.

Figure 4.5 A-D shows fluorescence emission spectra at different SH3 concentrations and

a constant H0 F10W concentration. With increasing SH3 concentration, quenching of Trp

emission is observed to increase. The quenching percentage of Trp emission was

calculated and plotted in figure 4.5E. The data were then fitted by a Langmuir isotherm

adsorption, which yielded a dissociation constant (kd) of 7.5 ± 1.9μΜ (uncertainty is from

standard error).

To exclude the possibility that quenching of Trp in SH3 was due to the introduction of

PheCN or Trp in H0 helix, I next tested the influence of H0 WT helix on the signal of Trp 68

in SH3. Figure 4.5F shows the quenching percentage of Trp in SH3 with presence of H0

WT. The quenching percentages induced by H0 WT are smaller than those induced by

H0 F10W, but still are mostly positive and increase with increasing SH3 concentration.

The Langmuir isotherm adsorption fitting results a dissociation constant (kd) of 4.5 ±

6.0μΜ (uncertainty is from standard error). The uncertainty is relative high, which is

probably due to the small quenching percentages.

D. Discussion

We investigated the autoinhibition mechanism of endophilin, by monitoring the changes

of fluorescence spectra in order to reveal the interactions between H0 and SH3. The

FRET experiments, which used H0 F10FCN as the donor and SH3 as the acceptor, led to

an interesting phenomenon of quenching of both donor and acceptor emission. We found

that this quenching happened with different donor to acceptor ratios and different donor

and acceptor concentrations. Furthermore, quenching increased with increasing

concentrations while the donor to acceptor ratio was held constant. Quenching of donor signal suggests that there is FRET between the H0 and SH3, which provides experimental evidences for binding between H0 and SH3.Though we did not observe sensitized emission, we believe this is due to simultaneous quenching of Trp, which is a known phenomenon of ligand binding to SH3 domains(1, 12). This quenching likely be so strong that it overcomes the sensitized emission.

Alternative to FRET experiments, we used Trp in the SH3 domain as a probe to detect the interaction between H0 and SH3. Both H0 F10W and H0 WT were proven to be able to induce quenching of the Trp in the SH3 domain. These quenching experiments further 69

prove binding between H0 and SH3 in solution. Both titrations yield similar dissociation constants, at the magnitude of ten micro molars. So far our result first reports the experimental determined binding affinity between separated H0 and SH3 of endophilin in solution. The simulation research which proposed this H0/SH3 interaction only reported a binding energy instead of a free energy of binding. The binding energy was found to be

13.0 ± 0.6 kcal/mol(9). A closer comparison might be the binding between endophilin

SH3 and PP19, which is a synaptoganin derived peptide, of which the dissociation constant was reported to be 14 μM (13). This result has similar magnitude compared to our results.

In conclusion, we found experimental evidences for the interaction between H0 and SH3 domain of endophilin through three different kinds of experiments, and obtained the binding affinity between them in solution. This research should greatly contribute to understanding of auto-inhibition mechanism of endophilin in solution, and elucidate the function of SH3 domain of endophilin which is a key player in regulating other accessory proteins in CME.

70

Figures

Figure 4.1: H0 F10FCN/ SH3 FRET.

A. Inspection of the equilibrated H0/ SH3 complex shows that the Trp327 and Phe338 residues surround the Phe10 residue, leading to a strong hydrophobic interaction. Panel A

is reproduced from(9) under permission from Publisher Elsevier Group.

B. The chemical structure of PheCN which is used to subsitute Phe10.

C. Steady-state emission spectruma of H0 F10FCN only (green hollow square), SH3 only

(red sphere, rescaled according to its excitation spectrum as used in the EmEx method(14)

(Briefly, the excitation spectra of the acceptor only and the acceptor with the donor were taken. The the emission spectrum of the acceptor would be rescaled according to the ratio

between the excitation spectra of the acceptor without or with the donor, before compared

with the emission spectrum of the acceptor and the donor)) and FRET between H0

F10FCN and SH3 (black line), respectively. H0 F10FCN was 300 μM and SH3 was 20 μM,

which yielded a donor to acceptor ratio of 15:1.

71

D. Presenting the result in panel C in a different format: the separated spectrum of H0

F10FCN only and SH3 only in (C). were summed and gave one spectrum (red hollow sphere). Excitation wavelengths are 240nm.

72

Figure 4.2: Determination of the linear regime of fluorescence intensity-concentration relationship for H0 F10FCN.

A. Emission spectra of H0 F10FCN at a series of concentrations.

B. Fluorescence intensity values at wavelength 295nm from emission spectrums in panel

A at different H0 F10FCN concentrations. The vertical black line shows the boundary of linear regime. Excitation wavelengths are 240nm.

73

Figure 4.3: Incubation of H0 F10FCN and SH3 leads to both donor and acceptor emission intensities decrease.

A-D. H0 F10FCN only (green hollow square), SH3 only (red sphere, rescaled according to its excitation spectrum as in the EmEx method(14)) . H0 F10FCN was mixed with SH3 domains (black line) at different solution concentrations but the same H0 F10FCN to SH3 ratio (2:1). Excitation wavelengths are 240nm.

74

Figure 4.4: Determine linear regime of fluorescence intensity-concentration

relationship for endophilin SH3 domain.

A. Emission spectra of SH3 domains at various concentrations.

B. Fluorescence intensities at wavelength 358nm from emission spectra in panel A at different SH3 solution concentrations. The vertical black line shows the boundary of the

linear regime. Excitation wavelengths are 280nm.

75

Figure 4.5: Titration of quenching of Trp in SH3 by H0 F10W or H0 WT.

A-D. Fluorescent emission spectra of SH3 domains at different concentrations with H0

F10W held constant at 400μM. The green line is the spectrum of incubated SH3 and H0

F10W, while the black line is the addition of separated SH3 and H0 F10W spectrums.

E. Fluorescence titration experiments for SH3 and H0 F10W. Black dots are data points

averaged from three independent experiments. Error bars are SEM. The data is fitted by

Langmuir adsorption isotherm adsorption shown as a black line.

76

F. Fluorescence titration experiments for SH3 and H0 WT. Black dots are data points from three independent experiments. Error bars are SEM. The data is fitted by Langmuir adsorption isotherm shown as a black line. Excitation wavelengths are 280nm.

77

E. References: 1. Ringstad, N., H. Gad, P. Löw, G. Di Paolo, L. Brodin, O. Shupliakov, and P. De Camilli. 1999. Endophilin/SH3p4 is required for the transition from early to late stages in clathrin-mediated synaptic vesicle endocytosis. Neuron 24:143-154. 2. Verstreken, P., O. Kjaerulff, T. E. Lloyd, R. Atkinson, Y. Zhou, I. A. Meinertzhagen, and H. J. Bellen. 2002. Endophilin mutations block clathrin-mediated endocytosis but not neurotransmitter release. Cell 109:101-112. 3. Llobet, A., J. L. Gallop, J. J. E. Burden, G. Camdere, P. Chandra, Y. Vallis, C. R. Hopkins, L. Lagnado, and H. T. McMahon. 2011. Endophilin drives the fast mode of vesicle retrieval in a ribbon synapse. The Journal of Neuroscience 31:8512-8519. 4. Gallop, J. L., C. C. Jao, H. M. Kent, P. J. G. Butler, P. R. Evans, R. Langen, and H. T. McMahon. 2006. Mechanism of endophilin N-BAR domain mediated membrane curvature. The EMBO Journal 25:2898-2910. 5. Ringstad, N., Y. Nemoto, and P. De Camilli. 1997. The SH3p4ySh3p8ySH3p13 protein family: Binding partners for synaptojanin and dynamin via a Grb2-like Src homology 3 domain. Proceedings of the National Academy of Science of the United States of America 94:8569–8574. 6. Sundborger, A., C. Soderblom, O. Vorontsova, E. Evergren, J. E. Hinshaw, and O. Shupliakov. 2011. An endophilin-dynamin complex promotes budding of clathrin-coated vesicles during synaptic vesicle recycling. Journal of Cell Science 124:133-143. 7. Wang, Q., H. Y. Kaan, R. N. Hooda, S. L. Goh, and H. Sondermann. 2008. Structure and plasticity of Endophilin and 9. Structure 16:1574-1587. 8. Rao, Y., Q. Ma, A. Vahedi-Faridi, A. Sundborger, A. Pechstein, D. Puchkov, L. Luo, O. Shupliakov, W. Saenger, and V. Haucke. 2010. Molecular basis for SH3 domain regulation of F-BAR-mediated membrane deformation. Proceedings of the National Academy of Science of the United States of America 107:8213-8218. 9. Va´zquez, F. X., V. M. Unger, and G. A. Voth. 2013. Autoinhibition of endophilin in solution via interdomain interactions. Biophysical Journal 104:396–403. 10. Capraro, B. R., Z. Shi, T. Wu, Z. Chen, J. M. Dunn, E. Rhoades, and T. Baumgart. 2013. Kinetics of endophilin N-BAR domain dimerization and membrane interactions. The Journal of Biological Chemistry 288:12533-12543. 11. Tucker, M. J., R. Oyola, and F. Gai. 2005. Conformational Distribution of a 14-Residue Peptide in Solution: A Fluorescence Resonance Energy Transfer Study. Journal of Physical Chemistry B 109:4788-4795. 12. Cowburn, D., J. Zheng, Q. Xu, and G. Barany. 1995. Enhanced afinities and specificities of consolidated ligands for the Src Homology (SH) 3 and SH2 Domains of abelson protein-tyrosine kinase. The Journal of Biological Chemistry 270:26738-26741. 13. Ringstad, N., Y. Nemoto, and P. De Camilli. 2001. Differential Expression of Endophilin 1 and 2 Dimers at Central Nervous System Synapses. The journal of biological chemistry 276:40424–40430. 78

14. Merzlyakov, M., L. Chen, and K. Hristova. 2007. Studies of receptor tyrosine kinase transmembrane domain interactions: the EmEx-FRET method. The Journal of Membrane Biology 215:93-103.

79

Chapter 5: Syndapin full length BAR/SH3 intramolecular autoinihbition

A. Background

Syndapin is an FBAR domain containing protein, which has three types: syndapin I, II

and III. Among them syndapin I is enriched at neuron(1), involved in CME and ADBE

(activity dependent bulk endocytosis) (2). More information about syndapin is discussed

in Chapter 1B. One mechanism that regulates the functionality of syndapin in CME is its autoinhibtion(3-4). It was reported that SH3/BAR domain clamping in syndapin inhibits its tubulation ability, and that this autoinhibition could be released by a proline rich domain (PRD) derived from dynamin(3). However, there are puzzles with this proposal: based on this hypothesis, the key residues in the BAR domain which interacts with the

SH3 domain, and where therefore are responsible for auto-inhibition, have also been found to be important for syndapin membrane binding (3, 5-6). However, no significant differences between the binding abilities of inhibited and released syndapin were found

(3). Furthermore, no mechanism is currently know that would explain why the dislodge of the SH3 domain would alter the membrane deformation ability of syndapin(4).

In order to find a possible mechanism, we performed membrane curvature sorting experiments of syndapin FBAR (SFB), syndapin full length (SFL) with or without proline rich domain peptide (P1), to compare their curvature sensing abilities. To understand inhibition of syndapin full length curvature generation capability, we carried out tether pulling force measurements. EM experiments were carried out to figure out the curvature initiation abilities of SFB and SFL with or without P1. We find that for curvature sensing and initiation abilities, SFL are significantly weaker than SFB; while the spontaneous curvatures of the membrane induced by these proteins are 80

indistinguishable. We did not observe significant a influence of P1 on SFL regarding any of its tubualtion related properties that we examined. Our results are the most quantitative and comprehensive analysis which breaks down the membrane deformation ability of syndapin to investigate the mechanism guiding its function. Based on our observations, we hypothesize that the lack of tubulation ability of syndapin in vivo, is shown as weak curvature initiation capacity in vitro, and this is mainly induced by insufficient membrane curvature sensing ability.

B. Materials and methods

1. Materials:

Proteins: recombinant mouse syndapin1 FBAR (1-337aa) and full length plasmids were generously provided by Pr. Haucke, Leibniz Institut für Molekulare Pharmakologie &

Freie Universität Berlin, Germany. Expression and purification of proteins were carried out as described in Appendix A.

Peptide: peptide P1 (3) was purchased from biological chemistry resource center (BCRC) at the Department of Chemistry, University of Pennsylvania.

Lipid: All synthetic lipids were purchased from Avanti. The lipid composition used for the sorting experiment and force experiment was DOPC 74%, DOPS 25%,

DSPE-Bio-PEG2000 0.7% and TR-DPHE 0.3%. The GUV formation is described in chapter 2B. The lipid composition for EM experiment was 100% DOPS. The LUVs were made by the standard technique of vesicle extrusion (7). 81

2. Methods:

Curvature sorting experiments were carried out as described in Chapter 2B.

Tether pulling force measurements: The tether pulling force measurements were

performed as described(8). The preparation of sample, measurement chamber and

micropipette were similar as those in the sorting experiments. The bead was controlled by

a custom optical tweezers(9) to pull a 20 µm tether from an aspirated vesicle. The membrane tension of the tether was changed and corresponding pulling force was recorded.

Negative Staining Electron Microscopy (EM): EM experiments were performed with assistance from our lab member Tingting Wu as described(10). Briefly, liposomes

(0.1mg/ml) were extruded as 400nm-diameter vesicles and incubated with or without proteins/ ligands at room temperature for 90 min. Then samples were applied to carbon/Formavar-supported copper grids (Electron Microscopy Sciences, Hatfield, PA) and incubated for 5 min. Then excess samples were removed from the grid by blotting on filter papers. Negative stain with 2% (w/v) uranyl acetate was applied on grids for 2 min followed by buffer washing and air dry. A JEM 1011 transmission electron microscope

(JEOL) was used to image the samples.

C. Results

82

1. Curvature sorting of syndapin FBAR/ full length with or without the PRD

domain.

To understand the mechanism of auto-inhibition of syndapin full length tubulation ability,

I first examined its curvature sensing ability. Fig. 5.1A shows curvature preference of

syndapin full length (the middle row): the merged image suggests the protein to lipid

signal ratio on the tether is higher than that on the GUV. This phenomenon indicates full

length syndapin is not entirely inhibited from curvature sensing. More quantitative

evidence is from the cross sectional images shown in fig. 5.1B: with the tension therefore

curvature increasing, the signal in the lipid channel on the tether decreased, while that in

the protein channel increased. Images as in fig. 5.1B were obtained at multiple tensions

and resulted in traces as shown in fig. 5.1C. The opposite tendencies of lipid and protein

signals on the tether at multiple tensions further prove that SFL is able to sense

membrane curvature.

Compared to the full length protein, syndapin FBAR (fig. 5.1A, the left row) behaves as a

stronger curvature sensor, which is indicated by the almost identical fluorescence

intensities on the tether and vesicle (Since for the molecules which are not sorting, their

intensities on the tether would be much more lower than those on the vesicles, due to the

property of the optical microscope). The presence of P1 (the proline rich domain derived

from dynamin 1(3)) seems to have little influence on the sensing ability of the full length

protein.

The sorting efficiencies of proteins with or without ligands are displayed in figs. 5.2.

Panel A-C show representative curvature-composition coupling relationships for

syndapin FBAR, full length and full length with P1, respectively. The results (fig. 5.2D) 83

again prove that full length syndapin possesses curvature sorting ability. However, it is

weaker compared to the FBAR domain (*P<0.05) and not significantly influenced by the

ligand.

2. Curvature generation of syndapin FBAR/ full length with or without the PRD

domain

I investigated the curvature generation abilities of syndapin FBAR and full length with or without P1 by monitoring the pulling forces on the tethers covered by these proteins

under multiple membrane tensions. For a bare lipid tether (fig. 5.3A), the force is

proportional to the square root of membrane tension ( ) and their linear regression passes through the original point. In the presence of syndapin√σ FBAR, the force still varied linearly with (fig. 5.3B). However, the linear regression results a non-zero intercept

with x ax √=σ 0.10 ± 0.03 (mN/m)1/2is, (uncertainty is from standard error). This ∗ nonzero intercept√σ means that syndapin FBAR induced spontaneous curvature of the

membrane C0, which indicates it has curvature generation ability(11). I then tested

syndapin full length curvature generation ability (fig. 5.3C). A non-zero intercept was also observed ( = 0.10 ± 0.01 (mN/m)1/2), which suggested that the full length ∗ protein retains its√σ curvature generation ability. The spontaneous curvatures induced by these proteins are proportional to the intercepts(12) therefore are indistinguishable, which

indicates that they have similar curvature generation abilities (T-test was conducted by

comparing between trials with SFBAR and SFL, resulted in P=0.55>0.05). The addition

of P1 has negligible influence on the full length protein since its linear regression also

84

results a similar non-zero intercept = 0.14 ± 0.02 (mN/m)1/2 (T-test was ∗ conducted by comparing between trials√ σwith SFL and SFL with P1, resulted in

P=0.07>0.05).

3. Curvature initiation of syndapin FBAR/ full length with or without the PRD

domain

The previous investigations were based on the membrane which contained a portion with

a high curvature. We then wanted to figure out how these proteins with or without ligands

transform the relatively flat membrane into tubules. We incubated the LUVs with

proteins with or without ligands and observed them under electron microscopy with

assistance from our lab member Tingting Wu. Figs. 5.4A-C show representative images of liposomes incubated with syndapin FBAR, full length, and full length with the ligand.

It is clear that syndapin FBAR is more efficient at curvature initiating compared to the full length protein, which shows low membrane tubulation ability. This observation could

be further quantitatively proved by comparing the radii of tubes generated by proteins

and ligands. The results in fig. 5.4D indicates the tubes generated by syndapin FBAR are

significantly narrower than syndapin full length with or without P1 (***P<0.001).

Meanwhile, the radii of tubes generated by full length syndapin with or without P1 are

indistinguishable.

D. Discussion

The low membrane deformation ability of syndapin full length has been indicated by both

in vivo and in vitro experiments (3), while the exact mechanism remains insufficiently 85

understood as already mentioned above. We broke down the membrane deformation

ability into membrane curvature sensing, generation and initiation capability, respectively;

then examined all of these abilities of syndapin FBAR, the full length protein and the full

length protein with the ligand P1. For the curvature initiation abilities of syndapin FBAR

and full length, we obtained similar results as reported before (3-4): syndapin FBAR

presents much stronger curvature initiation ability than the full length protein. We

hypothesize that this difference between syndapin FBAR and full length protein, is most

likely induced by the discrepancy of their membrane curvature sensing abilities, which

have been first examined and compared through our experiments. Besides, we did not

observe an obvious releasing effect of the full length protein through the presence of the

proline rich domain derived peptide P1. The insufficient influences of P1 on syndapin

curvature sensing and generation abilities, we are not very surprised, since we

hypothesize that the key step in this release process, is at the releasing of the curvature

initiation procedure, during the whole syndapin membrane deformation process. But we

did not observe a significant influence of P1 on curvature initiation of syndapin either

(through EM experiment). The possible explanation could be the order of incubating

different components actually is important in the in vitro experiments. Though we have

tried a few different orders and haven’t observed any differences, we haven’t exhausted

all the possible orders. Besides, the time interval between adding different components

could be varied. Also the peptide is highly charged and its theoretical PI is 12.48. Such sample could be easily precipitated in the neutral environment therefore compromise its functionality easily.

86

Figures

Figure 5.1: Curvature-dependent partition of syndapin FBAR (SFB) and syndapin

full length (SFL) on liposome.

A. Confocal fluorescence xy images of SFB/SFL (labeled with Alexa Fluor 488)

partitioning on the lipid membrane. The left, middle and right columns are for syndapin

FBAR, syndapin full length without and with P1, respectively. Scale bars: 5µm. The upper, middle and lower channels are merged images, images from protein channel and lipid channel, respectively. All syndapin FBAR, syndapin full length with or without P1

preferred to partition on the higher curvature tubular membrane (on the left) instead of

relatively flat membrane on the GUVs (right). These suggest both proteins could sense

membrane curvature.

B. Confocal cross section xz images of a membrane tether covered by syndapin full

length (as shown in panel A) at low/high tension. As the tension therefore membrane

curvature increased, the intensity signal in the lipid channel dropped, while that in the

protein channel increased.

C. A representative example of the quantification of intensities of both protein and lipid

87

channels from the cross section of a tether covered by syndapin full length at different tensions. Red circles are the data points for the lipid; while the green squares are the data points for the protein. Scale bars: 2µm.

88

Figure 5.2: Syndapin full length maintains curvature sorting ability, but weaker than syndapin FBAR.

A-C. Representative examples of syndapin FBAR/ full length/ full length with P1 curvature-composition coupling relationships. The black spheres are the data points and the black lines are linear fitting results.

D. Multiple results as A-C were binned and plotted. The white circle, black square and white triangle is for syndapin FBAR /full length /full length with P1, respectively. The gray and black error bars are standard deviation and standard error of mean, respectively.

The protein concentrations are 50 nM in solution.

89

Figure 5.3: Syndapin full length could generate the spontaneous curvature of the membrane, as well as the syndapin FBAR.

A-D. Representative force-curvature relationships on the tubular membranes pulled out

from the bare vesicles or covered by syndapin FBAR or syndapin full length with or

without P1, respectively. Data points are showed in black triangles and the error bars are

standard deviations. The black lines are linear fits.

E. The binned results from multiple traces as in A-D. Data points of bare lipid, lipid with syndapin FBAR, syndapin full length, and syndapin full length with P1 are in black triangle, black square, white circle and white diamond, respectively; and black dash line,

black dot line, black solid line and gray line are corresponding liner fitting. Error bars are

standard deviation. The protein solution concentrations are 1µM.

90

Figure 5.4: Syndapin full length is short of membrane curvature initiation ability.

A-C. Negative stain EM images. LUVs were incubated with syndapin FBAR, syndapin full length without or with P1, respectively.

D. Statistical analysis of the radii of the tubes generated by syndapin FBAR and full

length with or without P1. Scale bars: 500nm.

91

E. References: 1. Modregger, J., B. Ritter, B. Witter, M. Paulsson, and M. Plomann. 2000. All three PACSIN isoforms bind to endocytic proteins and inhibit endocytosis. Journal of Cell Science 113:4511-4521. 2. Clayton, E. L., V. Anggono, K. J. Smillie, N. Chau, P. J. Robinson, and M. A. Cousin. 2009. The phospho-dependent dynamin-syndapin interaction triggers activity-dependent bulk endocytosis of synaptic vesicles. The Journal of Neuroscience 29:7706–7717. 3. Rao, Y., Q. Ma, A. Vahedi-Faridi, A. Sundborger, A. Pechstein, D. Puchkov, L. Luo, O. Shupliakov, W. Saenger, and V. Haucke. 2010. Molecular basis for SH3 domain regulation of F-BAR-mediated membrane deformation. Proceedings of the National Academy of Science of the United States of America 107:8213-8218. 4. Goh, S. L., Q. Wang, L. J. Byrnes, and H. Sondermann. 2012. Versatile Membrane Deformation Potential of Activated Pacsin. Plos ONE 7:e51628. 5. Shimada, A., H. Niwa, K. Tsujita, S. Suetsugu, K. Nitta, K. Hanawa-Suetsugu, R. Akasaka, Y. Nishino, M. Toyama, L. Chen, Z. Liu, B. Wang, M. Yamamoto, T. Terada, A. Miyazawa, A. Tanaka, S. Sugano, M. Shirouzu, K. Nagayama, T. Takenawa, and S. Yokoyama. 2007. Curved EFC/F-BAR-domain dimers are joined end to end into a filament for membrane invagination in endocytosis. Cell 129:761-772. 6. Frost, A., R. Perera, A. Roux, K. Spasov, O. Destaing, E. H. Egelman, P. De Camilli, and V. M. Unger. 2008. Structural Basis of Membrane Invagination by F-BAR Domains. Cell 132:807-817. 7. MacDonald, R. C., R. I. MacDonald, B. P. M. Menco, K. Takeshita, N. K. Subbarao, and L.-r. Hu. 1991. Small-volume extrusion apparatus for preparation of large, unilamellar vesicles. Biochimica et Biophysica Acta 1061:291-303. 8. Heinrich, M., B. R. Capraro, A. Tian, J. M. Isas, R. Langen, and T. Baumgart. 2010. Quantifying membrane curvature generation of drosophila amphiphysin N-BAR domains. J. Phys. Chem. Lett. 1:3401. 9. Heinrich, M., A. Tian, C. Esposito, and T. Baumgart. 2010. Dynamic sorting of lipids and proteins in membrane tubes with a moving phase boundary Proceedings of the National Academy of Sciences of the United States of America 107:7208-7213. 10. Capraro, B. R., Z. Shi, T. Wu, Z. Chen, J. M. Dunn, E. Rhoades, and T. Baumgart. 2013. Kinetics of endophilin N-BAR domain dimerization and membrane interactions. The Journal of Biological Chemistry 288:12533-12543. 11. Sorre, B., A. Callan-Jones, J. Manzi, B. Goud, J. Prost, P. Bassereau, and A. Roux. 2011. Nature of curvature coupling of amphiphysin with membranes depends on its bound density. Proceedings of the National Academy of Sciences 109:173-178. 12. Sorre, B., A. Callan-Jones, J. Manzi, B. Goud, J. Prost, P. Bassereau, and A. Roux. 2012. Nature of curvature coupling of amphiphysin with membranes depends on its bound density. Proceedings of the National Academy of Science of the United States of America 109:173-178.

92

Chapter 6: Conclusions and outlook

A. Nonlinear sorting, curvature generation, and crowding of Endophilin NBAR on

tubular membranesa

We have experimentally characterized the curvature sorting of the NBAR domain of

endophilin A1. Consistent with earlier findings from our group, the NBAR domain is

observed to sense membrane curvature at low concentration, and to generate curvature at

higher concentrations(1-2). Our measurements reveal a sigmoidal curvature/composition

coupling isotherm and suggest that attractive protein/ protein interactions (implying

positive cooperativity) can be amplified through curvature/composition coupling. This

may imply that an ENBAR protein membrane coverage fraction regime exists where

small changes in membrane curvature (MC) lead to large changes in membrane coverage,

in a synergistic effect that increases the sensitivity of curvature sorting, as has previously

been suggested (16). We furthermore have developed an analytical model that captures

the observed sigmoidal curvature sorting and predicts the existence of a protein density

switch that may function to determine the fate of maturing endocytic membrane pits.

B. Membrane interacting helices influencing curvature sorting and membrane

kinetics of endophilin NBAR

We have quantitatively investigated the effects of N-terminal and insert helices of endophilin on its curvature sensing and membrane dissociation kinetics, by comparing

a This paragraph is reproduced from previously published work: Zhu, C., Das, S.L., and Baumgart, T. (2012).

Nonlinear sorting, curvature generation, and crowding of endophilin N-BAR on tubular membranes. Biophys. J. 102,

1837-1845.

93

endophilin NBAR WT with two helix deletion mutants. We found that for both the same protein membrane density and the same solution concentration, the helix deletion mutants were weaker curvature sensors compared to the WT protein. Therefore we conclude that both helices have important contributions to the endophilin NBAR curvature sorting. The relative importance of these two helices is indistinguishable since the sorting efficiencies of these two mutants do not have statistically significant differences. These findings will improve further understanding of the scaffolding and helix insertion mechanisms which drive the NBAR domain protein membrane deformation.

We have examined the membrane dissociation kinetics of the endophilin NBAR domain and its helix deletion mutant. Both of them behaved following a single exponential decay on the time-scale we monitored. We discovered that for both WT and the mutant, their membrane dissociation rates are positively correlated with their densities on the membrane. These phenomena suggest that both variants have the tendencies of oligomerization. We also found irreversible binding fractions for both variants, and the percentage of these irreversible binding components to the total membrane densities are constants at different protein membrane densities. We hypothesize that these remaining fractions might indicate higher order organization of proteins on the membrane(3) which might be an intriguing realm to look into in the future.

C. Endophilin full length H0/ SH3 intradimer / intermolecular autoinhibition

We investigated endophilin full length autoinihibition mechanism by testing the interaction between its H0 helix and SH3 domain. We first engaged a FRET pair

PheCN-Trp in separate H0 helix and SH3 domain molecules and obtained the steady state 94

FRET spectra. We found that incubation of the donor with the acceptor resulted in the

quenching of both donor and acceptor emission intensities. This simultaneous decrease

implied interactions between H0 helix and SH3 domain. We then tested quenching of the

trp in the SH3 domain in the presence of H0 F10W or H0 WT. We found that both

peptides induced quenching of the trp in the SH3 domain. We then conducted

fluorescence titration experiments and first obtained experimental binding affinities

between the peptides and SH3 domain. Our results shed light on the research of

autoinhibition of endophilin full length(4), which might be an important mechanism

involved in the regulation of the clathrin-mediated endocytosis (CME) process.

D. Syndapin full length BAR/ SH3 intramolecular autoinihbition

To understand the mechanism of autoinhibition of syndapin tubulation ability, we

investigated the curvature sensing, generation and initiation abilities of syndapin FBAR,

full length w/o proline rich domain P1. We found that syndapin FBAR was a stronger

curvature initiator than the full length protein, which was likely induced by its greater

curvature sensing ability. We observed that the proline rich ligand P1 has negligible

influence on the tubulation ability of the full length protein. For future experiments, the full length proline rich domain of dynamin might be a more efficient ligand(5) which could release the inhibited syndapin.

95

E. References:

1. Heinrich, M., B. R. Capraro, A. Tian, J. M. Isas, R. Langen, and T. Baumgart. 2010. Quantifying membrane curvature generation of drosophila amphiphysin N-BAR domains. J. Phys. Chem. Lett. 1:3401. 2. Baumgart, T., B. R. Capraro, C. Zhu, and S. L. Das. 2011. Thermodynamics and mechanics of membrane curvature generation and sensing by proteins and lipids. Ann. Rev. Phys. Chem. 62:483-506. 3. Simunovic, M., A. Srivastava, and G. A. Voth. 2013. Linear aggregation of proteins on the membrane as a prelude to membrane remodeling. Proceedings of the National Academy of Science of the United States of America. 4. Va´zquez, F. X., V. M. Unger, and G. A. Voth. 2013. Autoinhibition of endophilin in solution via interdomain interactions. Biophysical Journal 104:396–403. 5. Goh, S. L., Q. Wang, L. J. Byrnes, and H. Sondermann. 2012. Versatile Membrane Deformation Potential of Activated Pacsin. Plos ONE 7:e51628.

96

Appendix

A. Protein expression, purification and labeling

This chapter describes the general protein expression, purification and labeling

procedures for the proteins investigated in this thesis. There were mainly two kinds of

bacterial expressed recombinant proteins: Glutathione-S-transferase (GST) fusion proteins

and his-tagged proteins.

Protein expression:

Protein expression used BL21 E. coli were performed as described in our lab member Dr.

Benjamin R. Caparo’s dissertation. The cell pellets were collected through centrifugation and followed by ultrasonic lysis and centrifugation to obtain soluble fusion proteins in the lysate.

Protein purification:

GST fusion proteins:

All purification processes were performed through FPLC (fast protein liquid chromatography) in a 4 °C fridge. The lysate, which contained GST fusion proteins, was

applied to a GSTrap FF 5ml column (GE) multiple times. Then the fusion proteins were

eluted by GSH (reduced glutathione) buffer and collected. The fusion protein was then

digested by PreScission protease or thrombin at 4 deg or room temperature for couple of

hours. The digestion products were then applied to a Hitrap Q HP 5ml column (GE) or Hitrap

SP HP 5ml column (GE) for anion or cation exchange depending on their isoelectric points.

The fractions were collected and analyzed by SDS-PAGE gel.The selected fractions were

applied to a HiLoad 16/600 Superdex 200 pg column (size exclusion column, GE) to further

97

separate and purify the samples. Single fraction from size exclusion column elution was

collected, aliquoted, frozen and stored at -80°C.

His-tagged proteins:

All purification processes were performed through FPLC in at 4 °C. The lysate which contained his-tagged fusion proteins was applied to a Histrap HP 5ml column (GE) multiple times. Then the fusion proteins were eluted by imidazole buffer and collected. The following

processes were similar to those of GST fusion proteins.

Protein labeling:

This paragraph described protein labeling at their cystine positions with Alexa Fluor 488 C5

maleimide. Generally, the proteins were incubated with 2-10 fold molar excess of Alexa

Fluor 488 C5 maleimide at 4 °C for overnight while continuously rotating. 10 fold molar

excess of DTT (compared to the amount of dyes used) was added to stop the reaction. Then

the samples were applied to three connected Hitrap Desalting columns (GE) to separate the

dyes and labeled proteins.

98

B. Abbreviations:

AF-488: Alexa Fluor 488

BAR: Bin/Amphiphysin/Rvs

SH3: Src homology 3

CME: clathrin mediated endocytosis

DOPC: 1,2-dioleoyl-sn-glycero-3-phosphocholine

DOPG: 1,2-dioleoyl-sn-glycero-3-phospho-(1'-rac-glycerol)

DOPS: 1,2-dioleoyl-sn-glycero-3-phospho-L-serine

TR-DHPE: Texas Red® 1,2-dihexadecanoyl-sn-glycero-3-phosphoethanolamine, triethylammonium salt

DSPE-Bio-PEG2000:1,2-distearoyl-sn-glycero-3-phosphoethanolamine-N-[biotinyl(poly ethylene glycol)-2000] (ammonium salt)

TCEP: tris(2-carboxyethyl)phosphine

EM: electron microscopy

DTT: dithiothreitol

FPLC: fast protein liquid chromatography

FRAP: fluorescence recovery after photobleaching

GUV: giant unilamellar vesicles

LUV: large unilamellar vesicles

GST: Glutathione-S-transferase

SDS-PAGE: polyacrylamide gel electrophoresis in presence of sodium dodecyl sulfate

99

BIBLIOGRAPHY Arkhipov, A., Y. Yin, and K. Schulten. 2008. Four-scale description of membrane sculpting by BAR domains. Biophysical journal 95:2806-2821.

Arkhipov, A., Y. Yin, and K. Schulten. 2009. Membrane-bending mechanism of amphiphysin N-BAR domains. Biophysical journal 97:2727-2735.

Ayton, G. S., and G. A. Voth. 2010. Multiscale simulation of protein mediated membrane remodeling. Seminars in Cell & Developmental Biology 21:357-362.

Ayton, G. S., E. Lyman, V. Krishna, R. D. Swenson, C. Mim, V. M. Unger, and G. A.

Voth. 2009. New insights into BAR domain-induced membrane remodeling. Biophysical

Journal. 97:1616-1625.

Bai, X., G. Meng, M. Luo, and X. Zheng. 2012. Rigidity of wedge loop in PACSIN 3 protein Is a key factor in dictating diameters of tubules. The Journal of Biological

Chemistry 287:22387-22396.

Baumgart, T., B. R. Capraro, C. Zhu, and S. L. Das. 2011. Thermodynamics and

Mechanics of Membrane Curvature Generation and Sensing by Proteins and Lipids.

Annual Review of Physical Chemistry, Vol 62 62:483-506.

Bevington, P. R., and D. K. Robinson. 2003. Data reduction and error analysis for the physical sciences. McGraw-Hill, Boston.

Bhatia, V. K., K. L. Madsen, P. Bolinger, A. Kunding, P. Hedegard, U. Gether, and D.

Stamou. 2009. Amphipathic motifs in BAR domains are essential for membrane curvature sensing. The EMBO Journal 28:3303-3314.

Blood, P. D., and G. A. Voth. 2006. Direct observation of Bin/amphiphysin/Rvs (BAR) domain-induced membrane curvature by means of molecular dynamics simulations.

100

Proceedings of the National Academy of Sciences of the United States of America

103:15068-15072.

Blood, P. D., R. D. Swenson, and G. A. Voth. 2008. Factors influencing local membrane curvature induction by N-BAR domains as revealed by molecular dynamics simulations.

Biophysical journal 95:1866-1876.

Boucrot, E., A. Pick, G. Camdere, N. Liska, E. Evergren, H. T. McMahon, and M. M.

Kozlov. 2012. Membrane Fission Is Promoted by Insertion of Amphipathic Helices and

Is Restricted by Crescent BAR Domains. Cell 149:124-136.

Bukman, D. J., J. H. Yao, and M. Wortis. 1996. Stability of cylindrical vesicles under axial tension. Physical Review E 54:5463-5468.

Campelo, F., G. Fabrikant, H. T. McMahon, and M. M. Kozlov. 2010. Modeling membrane shaping by proteins: focus on EHD2 and N-BAR domains. FEBS Letters

584:1830-1839.

Campelo, F., H. T. McMahon, and M. M. Kozlov. 2008. The hydrophobic insertion mechanism of membrane curvature generation by proteins. Biophysical journal.

95:2325-2339.

Capraro, B. R., Y. Yoon, W. Cho, and T. Baumgart. 2010. Curvature sensing by the epsin

N-terminal homology (ENTH) domain measured on cylindrical lipid membrane tethers.

Journal of the American Chemical Society 132:1200 - 1201.

Capraro, B. R., Z. Shi, T. Wu, Z. Chen, J. M. Dunn, E. Rhoades, and T. Baumgart. 2013.

Kinetics of endophilin N-BAR domain dimerization and membrane interactions. The

Journal of Biological Chemistry 288:12533-12543.

101

Chen, Y.-J., P. Zhang, E. H. Egelman, and J. E. Hinshaw. 2004. The stalk region of

dynamin drives the constriction of dynamin tubes. Nature Structural & Molecular

Biology:574 - 575.

Clayton, E. L., V. Anggono, K. J. Smillie, N. Chau, P. J. Robinson, and M. A. Cousin.

2009. The phospho-dependent dynamin-syndapin interaction triggers activity-dependent bulk endocytosis of synaptic vesicles. The Journal of Neuroscience 29:7706–7717.

Cocucci, E., F. o. Aguet, S. Boulant, and T. Kirchhausen. 2012. The first five seconds in the life of a clathrin-coated pit. Cell 150:495–507.

Cowburn, D., J. Zheng, Q. Xu, and G. Barany. 1995. Enhanced afinities and specificities of consolidated ligands for the Src Homology (SH) 3 and SH2 Domains of abelson protein-tyrosine kinase. The Journal of Biological Chemistry 270:26738-26741.

Crank J. 2004. The Mathematics of Diffusion. Oxford University Press, New York.

Cui, H., G. S. Ayton, and G. A. Voth. 2009. Membrane binding by the endophilin

N-BAR domain. Biophysical Journal 97:2746-2753.

Dickman, D. K., J. A. Horne, I. A. Meinertzhagen, and T. L. Schwarz. 2005. A slowed classical pathway rather than kiss-and-run mediates endocytosis at synapses lacking synaptojanin and endophilin. Cell 123:521-533.

Doherty, G. J., and H. T. McMahon. 2009. Mechanisms of endocytosis. Annual Review of Biochemistry 78:857–902.

Edeling, M. A., S. Sanker, T. Shima, P. K. Umasankar, S. Höning, H. Y. Kim, L. A.

Davidson, S. C. Watkins, M. Tsang, D. J. Owen, and L. M. Traub. 2009. Structural requirements for PACSIN/Syndapin operation during zebrafish embryonic notochord development. Plos ONE 4:e8150. 102

Farsad, K., N. Ringstad, K. Takei, S. R. Floyd, K. Rose, and P. D. Camilli. 2001.

Generation of high curvature membranes mediated by direct endophilin bilayer

interactions. The Journal of Cell Biology. 155:193-200.

Fernandes, F., L. M. S. Loura, F. J. Chichon, J. L. Carrascosa, A. Fedorov, and M. Prieto.

2008. Role of helix 0 of the N-BAR domain in membrane curvature generation.

Biophysical Journal 94:3065-3073.

Ford, M. G. J., I. G. Mills, B. J. Peter, Y. Vallis, G. J. K. Praefcke, P. R. Evans, and H. T.

McMahon. 2002. Curvature of clathrin-coated pits driven by epsin. Nature 419:361-366.

Frost, A., R. Perera, A. Roux, K. Spasov, O. Destaing, E. H. Egelman, P. De Camilli, and

V. M. Unger. 2008. Structural Basis of Membrane Invagination by F-BAR Domains. Cell

132:807-817.

Frost, A., V. M. Unger, and P. De Camilli. 2009. The BAR domain superfamily: membrane-molding macromolecules. Cell 137:191-196.

Gallop, J. L., and H. T. McMahon. 2005. BAR domains and membrane curvature: bringing your curves to the BAR. Biochemical Society Symposia. 72:223-231.

Gallop, J. L., C. C. Jao, H. M. Kent, P. J. G. Butler, P. R. Evans, R. Langen, and H. T.

McMahon. 2006. Mechanism of endophilin N-BAR domain mediated membrane curvature. The EMBO Journal 25:2898-2910.

Goh, S. L., Q. Wang, L. J. Byrnes, and H. Sondermann. 2012. Versatile Membrane

Deformation Potential of Activated Pacsin. Plos ONE 7:e51628.

Hatzakis, N. S., V. K. Bhatia, J. Larsen, K. L. Madsen, P. Bolinger, A. H. Kunding, J.

Castillo, U. H. Gether, P., and D. Stamou. 2009. How curved membranes recruit amphipathic helices and protein anchoring motifs. Nature Cell Biology 5:835-841. 103

Heinrich, M., A. Tian, C. Esposito, and T. Baumgart. 2010. Dynamic sorting of lipids and

proteins in membrane tubes with a moving phase boundary. Proceedings of the National

Academy of Sciences of the United States of America 107:7208-7213.

Heinrich, M., B. R. Capraro, A. Tian, J. M. Isas, R. Langen, and T. Baumgart. 2010.

Quantifying membrane curvature generation of drosophila amphiphysin N-BAR domains.

The Journal of Physical Chemistry Letters. 1:3401-3406.

Henne, W. M., H. M. Kent, M. G. J. Ford, B. G. Hegde, O. Daumke, P. J. G. Butler, R.

Mittal, R. Langen, P. R. Evans, and H. T. McMahon. 2007. Structure and analysis of

FCHo2F-BAR domain: A dimerizing and membrane recruitment module that effects membrane curvature. Structure 15:839-852.

Higgins, M. K., and H. T. McMahon. 2002. Snap-shots of clathrin-mediated endocytosis.

Trends in Biochemical Sciences 27:257-263.

Hokanson, D. E., and E. M. Ostap. 2006. Myo1c binds tightly and specifically to phosphatidylinositol 4,5-bisphosphate and inositol 1,4,5-trisphosphate. Proceedings of the National Academy of Sciences of the United States of America. 103:3118-3123.

Jao, C. C., B. G. Hegde, J. L. Gallop, P. B. Hegde, H. T. McMahon, I. S. Haworth, and R.

Langen. 2010. Roles of Amphipathic Helices and the Bin/Amphiphysin/Rvs (BAR)

Domain of Endophilin in Membrane Curvature Generation. The Journal of Biological

Chemistry 285:20164-20170.

Kessels, M. M., and B. Qualmann. 2002. Syndapins integrate N-WASP in receptor-mediated endocytosis. The EMBO Journal 21:6083-6094.

Krauss, M., J.-Y. Jia, A. Roux, R. Beck, F. T. Wieland, D. C. Pietro, and V. Haucke.

2008. Arf1-gtp-induced tubule formation suggests a function of arf family proteins in 104

curvature acquisition at sites of vesicle budding. The Journal of Biological Chemistry

283:27717-27723.

Lee, M. C. S., L. Orci, S. Hamamoto, E. Futai, M. Ravazzola, and R. Schekman. 2005.

Sar1p N-terminal helix initiates membrane curvature and completes the fission of a

COPII vesicle. Cell 122:605-617.

Leibler S. 1986. Curvature Instability in Membranes. Journal de Physique 47:507 - 516.

Li, Y. H., R. Lipowsky, and R. Dimova. 2011. Membrane nanotubes induced by aqueous phase separation and stabilized by spontaneous curvature. Proceedings of the National

Academy of Sciences of the United States of America 108:4731-4736.

Lipowsky, R., and H. G. Dobereiner. 1998. Vesicles in contact with nanoparticles and colloids. Europhysics Letters 43:219-225.

Llobet, A., J. L. Gallop, J. J. E. Burden, G. Camdere, P. Chandra, Y. Vallis, C. R.

Hopkins, L. Lagnado, and H. T. McMahon. 2011. Endophilin drives the fast mode of vesicle retrieval in a ribbon synapse. The Journal of Neuroscience 31:8512-8519.

Llobet, A., J. L. Gallop, J. J. E. Burden, G. Camdere, P. Chandra, Y. Vallis, C. R.

Hopkins, L. Lagnado, and H. T. McMahon. 2011. Endophilin drives the fast mode of vesicle retrieval in a ribbon synapse. The Journal of Neuroscience 31:8512-8519.

Löw, C., U. Weininger, H. Lee, K. Schweimer, I. Neundorf, A. G. Beck-Sickinger, R. W.

Pastor, and J. Balbach. 2008. Structure and dynamics of helix-0 of the N-BAR domain in lipid micelles and bilayers. Biophysical Journal 95:4315-4323.

Lundmark, R., G. J. Doherty, Y. Vallis, B. J. Peter, and H. T. McMahon. 2008. Arf family GTP loading is activated by, and generates, positive membrane curvature.

Biochemical Journal 414:189-194. 105

MacDonald, R. C., R. I. MacDonald, B. P. M. Menco, K. Takeshita, N. K. Subbarao, and

L.-r. Hu. 1991. Small-volume extrusion apparatus for preparation of large, unilamellar vesicles. Biochimica et Biophysica Acta 1061:291-303.

Masuda, M., and N. Mochizuki. 2010. Structural characteristics of BAR domain

superfamily to sculpt the membrane. Seminars in Cell & Developmental Biology 21:

391–398.

Masuda, M., S. Takeda, M. Sone, T. Ohki, H. Mori, Y. Kamioka, and N. Mochizuki.

2006. Endophilin BAR domain drives membrane curvature by two newly identified structure-based mechanisms. The EMBO Journal 25:2889-2897.

Mathivet, L., S. Cribier, and P. F. Devaux. 1996. Shape Change and Physical Properties of Giant Phospholipid Vesicles Prepared in the Presence of an AC Electric Field.

Biophysical Journal 70:1112 - 1121.

McMahon, H. T., and E. Boucrot. 2011. Molecular mechanism and physiological functions of clathrin-mediated endocytosis. Nature Reviews Molecular Cell Biology

12:517-533

McMahon, H. T., and J. L. Gallop. 2005. Membrane curvature and mechanisms of dynamic cell membrane remodelling. Nature 438:590-596.

Merzlyakov, M., L. Chen, and K. Hristova. 2007. Studies of receptor tyrosine kinase transmembrane domain interactions: the EmEx-FRET method. The Journal of Membrane

Biology 215:93-103.

Mim, C., H. Cui, Joseph A. Gawronski-Salerno, A. Frost, E. Lyman, Gregory A. Voth, and Vinzenz M. Unger. 2012. Structural Basis of Membrane Bending by the N-BAR

Protein Endophilin. Cell 149:137-145. 106

Mizuno, N., C. C. Jao, R. Langen, and A. C. Steven. 2010. Multiple modes of

endophilin-mediated conversion of lipid vesicles into coated tubes: implications for

synaptic endocytosis. Journal of Biological Chemistry 285:23351-23358.

Modregger, J., B. Ritter, B. Witter, M. Paulsson, and M. Plomann. 2000. All three

PACSIN isoforms bind to endocytic proteins and inhibit endocytosis. Journal of Cell

Science 113:4511-4521.

Peter, B. J., H. M. Kent, I. G. Mills, Y. Vallis, P. J. G. Butler, P. R. Evans, and H. T.

McMahon. 2004. BAR domains as sensors of membrane curvature: the amphiphysin

BAR structure. Science 303:495-499.

Plomann, M., J. G. Wittmann, and M. G. Rudolph. 2010. A hinge in the distal end of the

PACSIN 2 F-BAR domain may contribute to membrane-curvature sensing. Journal of

Molecular Biology 400:129–136.

Praefcke, G. J. K., and H. T. McMahon. 2004. The dynamin superfamily: universal membrane tabulation and fission molecules? Nature Reviews Molecular Cell Biology

5:133-147.

Prinz, W. A., and J. E. Hinshaw. 2009. Membrane-bending proteins. Critical Reviews in

Biochemistry & Molecular Biology 44:278-291.

Rao, Y., Q. Ma, A. Vahedi-Faridi, A. Sundborger, A. Pechstein, D. Puchkov, L. Luo, O.

Shupliakov, W. Saenger, and V. Haucke. 2010. Molecular basis for SH3 domain regulation of F-BAR-mediated membrane deformation. Proceedings of the National

Academy of Science of the United States of America 107:8213-8218.

Rao, Y., Q. Ma, A. Vahedi-Faridi, A. Sundborger, A. Pechstein, D. Puchkov, L. Luo, O.

Shupliakov, W. Saenger, and V. Haucke. 2010. Molecular basis for SH3 domain 107

regulation of F-BAR-mediated membrane deformation. Proceedings of the National

Academy of Science of the United States of America 107:8213-8218.

Ringstad, N., H. Gad, P. Löw, G. Di Paolo, L. Brodin, O. Shupliakov, and P. De Camilli.

1999. Endophilin/SH3p4 is required for the transition from early to late stages in

clathrin-mediated synaptic vesicle endocytosis. Neuron 24:143-154.

Ringstad, N., Y. Nemoto, and P. De Camilli. 1997. The SH3p4ySh3p8ySH3p13 protein

family: Binding partners for synaptojanin and dynamin via a Grb2-like Src homology 3 domain. Proceedings of the National Academy of Science of the United States of

America 94:8569–8574.

Ringstad, N., Y. Nemoto, and P. De Camilli. 2001. Differential Expression of Endophilin

1 and 2 Dimers at Central Nervous System Synapses. The Journal of Bbiological

Chemistry 276:40424–40430.

Roux, A., G. Koster, M. Lenz, B. Sorre, J.-B. Manneville, P. Nassoy, and P. Bassereaua.

2010. Membrane curvature controls dynamin polymerization. Proceedings of the

National Academy of Science of the United States of America 107:4141-4146.

Schuske, K. R., J. E. Richmond, D. S. Matthies, W. S. Davis, S. Runz, D. A. Rube, A. M. van der Bliek, and E. M. Jorgensen. 2003. Endophilin is required for synaptic vesicle endocytosis by localizing synaptojanin. Neuron 40:749-762.

Seifert U., Berndl K., and Lipowsky R. 1991. Shape Transformation of Vesicles: Phase

Diagram for Spontaneous-curvature and bilayer-coupling models. Physical Review A

44:1182 - 1202.

Sens, P., and M. S. Turner. 2004. Theoretical Model for the Formation of Caveolae and

Similar Membrane Invaginations. Biophysical Journal 86:2049-2057. 108

Shimada, A., H. Niwa, K. Tsujita, S. Suetsugu, K. Nitta, K. Hanawa-Suetsugu, R.

Akasaka, Y. Nishino, M. Toyama, L. Chen, Z. Liu, B. Wang, M. Yamamoto, T. Terada,

A. Miyazawa, A. Tanaka, S. Sugano, M. Shirouzu, K. Nagayama, T. Takenawa, and S.

Yokoyama. 2007. Curved EFC/F-BAR-domain dimers are joined end to end into a filament for membrane invagination in endocytosis. Cell 129:761-772.

Simunovic, M., A. Srivastava, and G. A. Voth. 2013. Linear aggregation of proteins on

the membrane as a prelude to membrane remodeling. Proceedings of the National

Academy of Science of the United States of America 110: 20396-20401

Sorre, B., A. Callan-Jones, J. Manzi, B. Goud, J. Prost, P. Bassereau, and A. Roux. 2012.

Nature of curvature coupling of amphiphysin with membranes depends on its bound

density. Proceedings of the National Academy of Science of the United States of America

109:173-178.

Stachowiak, J. C., C. C. Hayden, and D. Y. Sasaki. 2010. Steric confinement of proteins

on lipid membranes can drive curvature and tubulation. Proceedings of the National

Academy of Science of the United States of America 107:7781–7786.

Sundborger, A., C. Soderblom, O. Vorontsova, E. Evergren, J. E. Hinshaw, and O.

Shupliakov. 2011. An endophilin-dynamin complex promotes budding of clathrin-coated

vesicles during synaptic vesicle recycling. Journal of Cell Science 124:133-143.

Takei, K., V. I. Slepnev, V. Haucke, and P. D. Camilli. 1999. Functional partnership

between amphiphysin and dynamin in clathrinmediated endocytosis. Nature Cell Biology

1:33-39.

109

Taylor, M. J., D. Perrais, and C. J. Merrifield. 2011. A High Precision Survey of the

Molecular Dynamics of Mammalian Clathrin-Mediated Endocytosis. Plos Biology

9:e1000604.

Tian, A., and T. Baumgart. 2009. Sorting of lipids and proteins in membrane curvature

gradients. Biophysical Journal 96:2676 - 2688.

Tian, A., B. R. Capraro, C. Esposito, and T. Baumgart. 2009. Bending stiffness depends

on curvature of ternary lipid mixture tubular membranes. Biophysical Journal

97:1636-1646.

Tucker, M. J., R. Oyola, and F. Gai. 2005. Conformational Distribution of a 14-Residue

Peptide in Solution: A Fluorescence Resonance Energy Transfer Study. Journal of

Physical Chemistry B 109:4788-4795.

Va´zquez, F. X., V. M. Unger, and G. A. Voth. 2013. Autoinhibition of endophilin in solution via interdomain interactions. Biophysical Journal 104:396–403.

Verstreken, P., O. Kjaerulff, T. E. Lloyd, R. Atkinson, Y. Zhou, I. A. Meinertzhagen, and

H. J. Bellen. 2002. Endophilin mutations block clathrin-mediated endocytosis but not neurotransmitter release. Cell 109:101-112.

Wang, Q., H. Y. Kaan, R. N. Hooda, S. L. Goh, and H. Sondermann. 2008. Structure and plasticity of Endophilin and Sorting Nexin 9. Structure 16:1574-1587.

Wang, Q., M. V. A. S. Navarro, G. Peng, E. Molinelli, S. Lin Goh, B. L. Judson, K. R.

Rajashankar, and H. Sondermann. 2009. Molecular mechanism of membrane constriction

and tubulation mediated by the F-BAR protein Pacsin/Syndapin. Proceedings of the

National Academy of Sciences of the United States of America 106: 12700-12705

110

Watanabe, S., B. R. Rost, M. Camacho-Pe´rez, M. W. Davis, B. So¨hl-Kielczynski, C.

Rosenmund, and E. M. Jorgensen. 2013. Ultrafast endocytosis at mouse hippocampal

synapses. Nature 504:242–247.

Weissenhorn, W. 2005. Crystal structure of the Endophilin-A1 BAR domain. Journal of

Molecular Biology 35:653-661.

Yoon, Y., J. Tong, P. J. Lee, A. Albanese, N. Bhardwaj, M. Källberg, M. A. Digman, H.

Lu, E. Gratton, Y.-K. Shin, and W. Cho. 2010. Molecular basis of the potent

membrane-remodeling activity of the epsin 1 N-terminal homology domain. Journal of

Biological Chemistry 285:531-540.

Zhang, P., and J. E. Hinshaw. 2001. Three-dimensional reconstruction of dynamin in the

constricted state. Nature Cell Biology 3:922 - 926.

Zhu, C., S. L. Das, and T. Baumgart. 2012. Nonlinear sorting, curvature generation, and crowding of endophilin N-BAR on tubular membranes. Biophysical Journal

102:1837-1845.

Zimmerberg, J. 2004. Membrane curvature: How BAR domains bend bilayers Current

Biology 14:R250-R252.

Zimmerberg, J., and M. M. Kozlov. 2006. How proteins produce cellular membrane

curvature. Nature Reviews Molecular Cell Biology 7:9-19.

111