introduction to md simulations of dna systems › training › workshop › urbana2018c › ... ·...

48
Introduction to MD simulations of DNA systems Aleksei Aksimentiev

Upload: others

Post on 27-Jun-2020

3 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Introduction to MD simulations of DNA systems › Training › Workshop › Urbana2018c › ... · 2019-02-05 · 1 2 4 6 10 2 4 6 100 ar) 22 24 26 28 30 32 DNA-DNA distance (Å)

Introduction to MD simulations of DNA systems

Aleksei Aksimentiev

Page 2: Introduction to MD simulations of DNA systems › Training › Workshop › Urbana2018c › ... · 2019-02-05 · 1 2 4 6 10 2 4 6 100 ar) 22 24 26 28 30 32 DNA-DNA distance (Å)

Biological Modeling at Different Scalesspanning orders of magnitude in space and time

Page 3: Introduction to MD simulations of DNA systems › Training › Workshop › Urbana2018c › ... · 2019-02-05 · 1 2 4 6 10 2 4 6 100 ar) 22 24 26 28 30 32 DNA-DNA distance (Å)

2.5 nm

The computational microscopeMassive parallel computer

Blue Waters (UIUC): ~200,000 CPUs

Time scale: ~ 0.1-100 µs

Length scale: 10K - 1,000Matoms or (< 70 nm)3

Time resolution: 2 fsSpacial resolution: 0.1 A

Atoms move according to classical mechanics (F= ma)

Interactions between atoms are defined bythe molecular force field

Page 4: Introduction to MD simulations of DNA systems › Training › Workshop › Urbana2018c › ... · 2019-02-05 · 1 2 4 6 10 2 4 6 100 ar) 22 24 26 28 30 32 DNA-DNA distance (Å)

What is Molecular Dynamics?

Energy function has two parts:

- chemical bond interactions

- non-bonded interactions

Page 5: Introduction to MD simulations of DNA systems › Training › Workshop › Urbana2018c › ... · 2019-02-05 · 1 2 4 6 10 2 4 6 100 ar) 22 24 26 28 30 32 DNA-DNA distance (Å)

Chemical StructureBonds: Every pair of covalently bonded atoms.

Angles: Two bonds that share a common atom form an angle.

Dihedrals: Two angles that share a common bond form a dihedral.

Impropers: Any planar group of four atoms forms an improper.

Page 6: Introduction to MD simulations of DNA systems › Training › Workshop › Urbana2018c › ... · 2019-02-05 · 1 2 4 6 10 2 4 6 100 ar) 22 24 26 28 30 32 DNA-DNA distance (Å)

bond

angle

dihedral

Lennard-Jones (van der Waals)

electrostatic

The MD method models atoms withpoint par3cles interac3ng throughclassicalpoten3als

Page 7: Introduction to MD simulations of DNA systems › Training › Workshop › Urbana2018c › ... · 2019-02-05 · 1 2 4 6 10 2 4 6 100 ar) 22 24 26 28 30 32 DNA-DNA distance (Å)

MD force field: a approach

- Parametrize parts of the structure

- Assemble the parts

Example: a protein is a collection of 20 amino acids

Bonding cannot change(!)

Page 8: Introduction to MD simulations of DNA systems › Training › Workshop › Urbana2018c › ... · 2019-02-05 · 1 2 4 6 10 2 4 6 100 ar) 22 24 26 28 30 32 DNA-DNA distance (Å)

To run MD you’ll need…

"8

Computer (multi-core/GPU or cluster)

MD code (NAMD, Gromacs, Charmm, …)

Input coordinates (typically .pdb): specifies initial state of the system

Protein structure file (typically .psf): specifies chemical bond information

Parameter file: provides atom-type specific values to interacting potentials

Configuration file (.namd or similar): run-type parameter, temperature, etc

Constraints or other simulation-specific files (typically .pdb)

Output: MD trajectory (.dcd file): describes how coordinates change in time

Analysis tools (VMD, python, etc.

Page 9: Introduction to MD simulations of DNA systems › Training › Workshop › Urbana2018c › ... · 2019-02-05 · 1 2 4 6 10 2 4 6 100 ar) 22 24 26 28 30 32 DNA-DNA distance (Å)

What is MD good for?Minimumrequirements: -Well-definedques3on -Atomic-detailstructure(ordisorderedstate) -Minimumchemistry -Volumeof60x60x60nm3orless -Processestakinglessthan100microseconds

Typicalques3ons: -Whatistheequilibriumstructureinphysiologicalenvironment? -Whatisthefree-energycostofgoingfromAtoB? -WhatisthepathfromAandB? -Howdoesthestructurerespondtoexternalforceorfield?

-Whatisgoingoninmyexperiment?

Page 10: Introduction to MD simulations of DNA systems › Training › Workshop › Urbana2018c › ... · 2019-02-05 · 1 2 4 6 10 2 4 6 100 ar) 22 24 26 28 30 32 DNA-DNA distance (Å)

All-Atom Molecular Dynamics Simulation of DNA Condensates

Add 64 DNA helices

Add polyamine cations (+4)

Add 150 mM NaCl

Add explicit water

Solve the equation of motion (F= ma) under periodic

boundary condition in all directions

DNA-confining wall of radius RApply a half-harmonic wall

potential only to DNA

Page 11: Introduction to MD simulations of DNA systems › Training › Workshop › Urbana2018c › ... · 2019-02-05 · 1 2 4 6 10 2 4 6 100 ar) 22 24 26 28 30 32 DNA-DNA distance (Å)

1

2

46

10

2

46

100

Pres

sure

(ba

r)323028262422

DNA-DNA distance (Å)

Rau et al. AMBER99+CUFIX AMBER99 CHARMM

1

2

46

10

2

46

100

Pres

sure

(ba

r)

3025DNA-DNA distance (Å)

Rau et al. AMBER99+CUFIX AMBER99 CHARMM

CUFIX Improves Simulations of DNA Condensates

-120

-80

-40

0

40

80

120

y (Å

)

-120 -80 -40 0 40 80 120x (Å)

3

2

1

0

Charge conc. (M)

3210M

-120

-80

-40

0

40

80

120

y (Å

)

-120 -80 -40 0 40 80 120x (Å)

3

2

1

0

M

3210M

-120

-80

-40

0

40

80

120

y (Å

)-120 -80 -40 0 40 80 120

x (Å)

3

2

1

0

M

[Na] = 250 mM [Mg] = 20 mM [spermine] = 2 mM

AMBER99 AMBER99

2

4

12

4

102

4

100

Pres

sure

(ba

r)

3028262422DNA-DNA distance (Å)

Todd et al. AMBER99+CUFIX AMBER99 CHARMM

AMBER99

AMBER99 + CUFIX

AMBER99 + CUFIX

Yoo & Aksimentiev, NAR 2016

AMBER99 + CUFIX (MD included 200 mM Na)

CHARMM27

CHARMM27

CHARMM27

http://bionano.physics.illinois.edu/CUFIX

Page 12: Introduction to MD simulations of DNA systems › Training › Workshop › Urbana2018c › ... · 2019-02-05 · 1 2 4 6 10 2 4 6 100 ar) 22 24 26 28 30 32 DNA-DNA distance (Å)

C

D

D

A

crossover plane ii crossover plane iii

y

z

x

i ii iii i ii iii

12

34

56

1314

1516

1718

1211

109

87

B

y

x

crossover plane i

18

17

16

15

14

13

12

11

10

9

8

7

6

5

4

3

2

1

E

i ii iii i ii iii i ii iii i

�12

From caDNAno to all-atom

JejoongYoo

✴ CHARMM36 force field✴ Explicit water✴ [MgCl2] ~ 10 mM✴ NAMD✴ 1 to 3M atoms✴ 500 to 1,000 CPUs

Cadnano designIdeal structureSolvated structure

SM. Douglas,…, WM. Shih, Nucleic Acids Res., 2009

caDNAno2pdb

Page 13: Introduction to MD simulations of DNA systems › Training › Workshop › Urbana2018c › ... · 2019-02-05 · 1 2 4 6 10 2 4 6 100 ar) 22 24 26 28 30 32 DNA-DNA distance (Å)

Structural dynamics

Page 14: Introduction to MD simulations of DNA systems › Training › Workshop › Urbana2018c › ... · 2019-02-05 · 1 2 4 6 10 2 4 6 100 ar) 22 24 26 28 30 32 DNA-DNA distance (Å)

-2

-1

0

1

2

Tors

ion

per A

rray

Cel

l (°)

24.224.023.823.623.4Length per Array Cell (Å)

30

28

26

24

22

20

18

DNA-DNA distance (Å)- Inter-DNA distance in color map- Chicken wire frame represents

center line of helices & junction

Yoo and Aksimentiev, PNAS 110:20099 (2013)

Structural fluctuation of DNA origami nanorod

Cross-section

Structural fluctuations reveal local mechanical properties

80

60

40

20

0

Pers

isten

ce le

ngth

(µm

)

HC0 SQ0

perslp

µm

Persistence length of DNA origami

Simulations predict higher rigidity for honeycomb-lattice

design

Page 15: Introduction to MD simulations of DNA systems › Training › Workshop › Urbana2018c › ... · 2019-02-05 · 1 2 4 6 10 2 4 6 100 ar) 22 24 26 28 30 32 DNA-DNA distance (Å)

A Practical Guide to DNA Origami Simulations Using NAMD

Cadnano design

All-atom model

After optimization

cadnano2pdb web server

All-atom structure in a MgCl2 solution

— Walk through the protocol for all-atom simulations of DNA origami using the NAMD package

Perform MD simulation and analysis

Page 16: Introduction to MD simulations of DNA systems › Training › Workshop › Urbana2018c › ... · 2019-02-05 · 1 2 4 6 10 2 4 6 100 ar) 22 24 26 28 30 32 DNA-DNA distance (Å)

Chen-Yu Li

Page 17: Introduction to MD simulations of DNA systems › Training › Workshop › Urbana2018c › ... · 2019-02-05 · 1 2 4 6 10 2 4 6 100 ar) 22 24 26 28 30 32 DNA-DNA distance (Å)

Tiled DNA nanostructures

DNAbricks

DNAOrigami

Ke et al. Science 338:1177 (2012)

Origami BricksIdentical nucleotide sequence

in both structures

All-atom MD165 ns each

nanohub.org/resources/legogenSlone et al., New J. Phys. 18:055012 (2016)

Page 18: Introduction to MD simulations of DNA systems › Training › Workshop › Urbana2018c › ... · 2019-02-05 · 1 2 4 6 10 2 4 6 100 ar) 22 24 26 28 30 32 DNA-DNA distance (Å)

ENRG MD tool kit : a universal all-atom structure converter

CaDNAnano

NAB

CanDo

oxDNA

DAEDALUS

Nanoengineer vHelix

Tiamat

NAMD all-atom MD Simulation Files hextube.psfhextube.pdb

hextube.namdhexube.exb

CHARMM parameters

enrgmd tool-kit#enrgmd hextube.json

Page 19: Introduction to MD simulations of DNA systems › Training › Workshop › Urbana2018c › ... · 2019-02-05 · 1 2 4 6 10 2 4 6 100 ar) 22 24 26 28 30 32 DNA-DNA distance (Å)

Petascale computer systemHigh-resolution cryo-electron microscop

Cryo-EM reconstruction versus all-atom simulation

Page 20: Introduction to MD simulations of DNA systems › Training › Workshop › Urbana2018c › ... · 2019-02-05 · 1 2 4 6 10 2 4 6 100 ar) 22 24 26 28 30 32 DNA-DNA distance (Å)

Cryo-EM reconstruction versus all-atom simulation

Bai et al, PNAS 109:20012 (2012)

Page 21: Introduction to MD simulations of DNA systems › Training › Workshop › Urbana2018c › ... · 2019-02-05 · 1 2 4 6 10 2 4 6 100 ar) 22 24 26 28 30 32 DNA-DNA distance (Å)

Cryo-EM reconstruction versus all-atom simulation

Bai et al, PNAS 109:20012 (2012)

Page 22: Introduction to MD simulations of DNA systems › Training › Workshop › Urbana2018c › ... · 2019-02-05 · 1 2 4 6 10 2 4 6 100 ar) 22 24 26 28 30 32 DNA-DNA distance (Å)

Cryo-EM reconstruction versus all-atom simulation

Bai et al, PNAS 109:20012 (2012)

Pseudo-atomic model

Page 23: Introduction to MD simulations of DNA systems › Training › Workshop › Urbana2018c › ... · 2019-02-05 · 1 2 4 6 10 2 4 6 100 ar) 22 24 26 28 30 32 DNA-DNA distance (Å)

MD simulation of the cryo-EM object starting from a caDNAno design

7M atom solvated model~200 ns MD trajectory

Bai et al, PNAS 109:20012 (2012)

Page 24: Introduction to MD simulations of DNA systems › Training › Workshop › Urbana2018c › ... · 2019-02-05 · 1 2 4 6 10 2 4 6 100 ar) 22 24 26 28 30 32 DNA-DNA distance (Å)

7M atom solvated model~200 ns MD trajectory

Bai et al, PNAS 109:20012 (2012)

MD simulation of the cryo-EM object starting from a caDNAno design

Page 25: Introduction to MD simulations of DNA systems › Training › Workshop › Urbana2018c › ... · 2019-02-05 · 1 2 4 6 10 2 4 6 100 ar) 22 24 26 28 30 32 DNA-DNA distance (Å)

7M atom solvated model~200 ns MD trajectory

Bai et al, PNAS 109:20012 (2012)

MD simulation of the cryo-EM object starting from a caDNAno design

Page 26: Introduction to MD simulations of DNA systems › Training › Workshop › Urbana2018c › ... · 2019-02-05 · 1 2 4 6 10 2 4 6 100 ar) 22 24 26 28 30 32 DNA-DNA distance (Å)

Electron density maps

Cryo-EM reconstruction All-atom MD simulation

Page 27: Introduction to MD simulations of DNA systems › Training › Workshop › Urbana2018c › ... · 2019-02-05 · 1 2 4 6 10 2 4 6 100 ar) 22 24 26 28 30 32 DNA-DNA distance (Å)

Comparison with experiment

simulationEM density psuedo-atomic model

Maffeo, Yoo & Aksimentiev, NAR 44: 3013 (2016)

Page 28: Introduction to MD simulations of DNA systems › Training › Workshop › Urbana2018c › ... · 2019-02-05 · 1 2 4 6 10 2 4 6 100 ar) 22 24 26 28 30 32 DNA-DNA distance (Å)

Making images and animations of DNA nanostructures with VMD

Images from a

single structure

Image with multiple

representations

Animation based

on a trajectory

Animation including

transitions

Page 29: Introduction to MD simulations of DNA systems › Training › Workshop › Urbana2018c › ... · 2019-02-05 · 1 2 4 6 10 2 4 6 100 ar) 22 24 26 28 30 32 DNA-DNA distance (Å)

Elastic network of restraints guided MD (ENRG MD)

Solvent replaced with elastic

network

~10,000 times more efficient

Maffeo, Yoo & Aksimentiev, NAR 44: 3013 (2016)

Page 30: Introduction to MD simulations of DNA systems › Training › Workshop › Urbana2018c › ... · 2019-02-05 · 1 2 4 6 10 2 4 6 100 ar) 22 24 26 28 30 32 DNA-DNA distance (Å)
Page 31: Introduction to MD simulations of DNA systems › Training › Workshop › Urbana2018c › ... · 2019-02-05 · 1 2 4 6 10 2 4 6 100 ar) 22 24 26 28 30 32 DNA-DNA distance (Å)

Curved DNA origami systems

10

8

6

4

2

0

Bend

per

7-b

p ar

ray

cell

(°)

24201612840Array cell index

Programmed bend

120

110

100

90

80

70

60En

d-en

d an

gle

(°)

302520151050Time (ns)18

17

16

15

14

13

12

11

10

9

8

7

6

5

4

3

2

1

t̂1t̂2

t̂3

t̂3

t̂3

234567891011

1213

14

15

16

17

18

19

20

21

22

23

24

25

s = L

s = 0

1

26

27

Programm

ed

bend

Experiment Dietz, H. et al, Science, 325

Yoo and Aksimentiev, PNAS 110:20099 (2013)

Page 32: Introduction to MD simulations of DNA systems › Training › Workshop › Urbana2018c › ... · 2019-02-05 · 1 2 4 6 10 2 4 6 100 ar) 22 24 26 28 30 32 DNA-DNA distance (Å)

Multi-resolution simulations of self-assembled DNA nanostructures

Page 33: Introduction to MD simulations of DNA systems › Training › Workshop › Urbana2018c › ... · 2019-02-05 · 1 2 4 6 10 2 4 6 100 ar) 22 24 26 28 30 32 DNA-DNA distance (Å)

DNA membrane channelsDr.UlrichF.KeyserCambridge,UK

Page 34: Introduction to MD simulations of DNA systems › Training › Workshop › Urbana2018c › ... · 2019-02-05 · 1 2 4 6 10 2 4 6 100 ar) 22 24 26 28 30 32 DNA-DNA distance (Å)

DNA Ion Channels

previously published nanopore,[10] but the current design isconsiderably simpler, as six rather than 14 strands are used.Crucially, two porphyrin tags are positioned at the end of thenanobarrel to achieve its insertion into the bilayer in a direc-tional manner (Figure 1b). To facilitate the comparison to ourprevious nanopore work, we used a design with two DNAstrands that contain phosphorothiate groups in place of thephosphates in a number of positions in the backbone. Thethioate groups behave like the native negatively chargedphosphate groups at our experimental conditions of pH 8.0(Table S1).[10]

The DNA nanopore was assembled by heating andcooling an equimolar mixture of four regular and two TPP-modified DNA strands (for sequences see Table S1). Theassembly mixture was characterized to confirm the correctand successful formation of the DNA nanopore. Native gelelectrophoresis yielded a band, which migrated to the sameheight as a control nanopore without the porphyrin anchor(Figure 2a, lanes 2 and 3, respectively; main band co-migrat-ing at the 550 bp marker). The tailing of the band for theporphyrin DNA pore does not indicate unfolding, but israther caused by the very hydrophobic tag.[9,19,20] Theconcerted assembly into the nanobarrels was also confirmedby a single defined transition in the UV melting profiles,(Figure 2b; Tm= 53.4! 1.0 8C; n= 3); the opposite and unex-pected independent hybridization of the multicomponentDNA duplexes of different melting temperatures would haveled to a very broad transition. Furthermore, dynamic lightscattering (DLS) established the monomeric nature of thenanobarrels, as only a single peak with a hydrodynamic radiusof 5.5! 0.1 nm was observed (Figure 2c). The radius is largerthan the calculated value of 4.9 nm,[22,23] but in line with theaccuracy of DLS measurements for related DNA nano-structures.[23,24] The detailed dimensions of the nanobarrelwere established with atomic force microscopy (AFM)analysis (Figure 2d). The apparent height of (2.20!0.25) nm was expected for tip-compressed hollow DNAnanostructures.[8,25] Similarly, the AFM-derived length and

width of (20.4! 4.5) nm and (9.7! 2) nm (full width at halfmaximum), respectively, were in good agreement with thetheoretical dimensions (14 nm and 5.5 nm) after correctingfor tip-deconvolution.[10,26]

Having completed their structural characterization, weinvestigated whether the porphyrin-tagged DNA nanoporecan be stably anchored into lipid bilayers. In this examinationwe took advantage of the fluorescence properties of theporphyrin tag. Once incubated with giant unilamellar vesicles,the tagged nanopores could be visualized using fluorescencemicroscopy (Figure 3a, bottom). A bright-field image docu-mented the shape of the vesicle (Figure 3a, top). The brightfluorescence spots in the ring-shaped bilayer may representindividual DNA nanopores, given that the imaging conditionsof our custom-built set-up are powerful enough to detectsingle fluorophores.[27] However, as the DNA pores rapidlydiffuse out of the focal plane (Supporting Information,Movie SM1) and exhibit surprising photostability, it is hardto detect individual pores by single-fluorophore photobleach-ing. The anchoring of the nanopore into the bilayer wasconfirmed by independent spectroscopic analysis based on

Figure 1. DNA-nanopore-carrying porphyrin-based lipid anchors.a) Deoxyuridine bonded to tetraphenylporphyrin (TPP) through anacetylene linkage at the 5 position of the nucleobase. b) A DNAnanopore composed of six interconnected duplexes, represented ascylinders. The six-component DNA oligonucleotides are shown ingreen. The magenta porphyrin tags anchor the DNA nanopore into thelipid bilayer. For reasons of visual clarity, only the porphyrin corewithout the phenyl groups is shown. The TPP tags are not drawn toscale.

Figure 2. Characterization of porphyrin-modified nanobarrels assem-bled from DNA oligonucleotides. a) Agarose gel electrophoresis. Lane1, 100 bp marker; lane 2, DNA nanopore; lane 3, TPP-DNA nanopore.b) UV melting profile of a TPP nanopore. c) Dynamic light scatteringtrace of TPP nanopore. d) AFM micrograph of individual nanopores(arrows), and of elongated nanopore assemblies, which are stabilizedby inter-barrel base stacking; scale bar=50 nm. abs.=absorbance,T= temperature.

Figure 3. Porphyrin tags anchor DNA nanopores into lipid bilayers.a) Microscopic bright-field (top) and fluorescence (bottom;lexc=532 nm) images of a DPhPC vesicle containing DNA nanopores;scale bars=5 mm. b) The fluorescence emission spectrum for the TPPnanopore (c) is shifted upon insertion into vesicle bilayers (a ;lexc=424 nm). F= fluorescence intensity.

.AngewandteCommunications

12070 www.angewandte.org ! 2013 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim Angew. Chem. Int. Ed. 2013, 52, 12069 –12072

previously published nanopore,[10] but the current design isconsiderably simpler, as six rather than 14 strands are used.Crucially, two porphyrin tags are positioned at the end of thenanobarrel to achieve its insertion into the bilayer in a direc-tional manner (Figure 1b). To facilitate the comparison to ourprevious nanopore work, we used a design with two DNAstrands that contain phosphorothiate groups in place of thephosphates in a number of positions in the backbone. Thethioate groups behave like the native negatively chargedphosphate groups at our experimental conditions of pH 8.0(Table S1).[10]

The DNA nanopore was assembled by heating andcooling an equimolar mixture of four regular and two TPP-modified DNA strands (for sequences see Table S1). Theassembly mixture was characterized to confirm the correctand successful formation of the DNA nanopore. Native gelelectrophoresis yielded a band, which migrated to the sameheight as a control nanopore without the porphyrin anchor(Figure 2a, lanes 2 and 3, respectively; main band co-migrat-ing at the 550 bp marker). The tailing of the band for theporphyrin DNA pore does not indicate unfolding, but israther caused by the very hydrophobic tag.[9,19,20] Theconcerted assembly into the nanobarrels was also confirmedby a single defined transition in the UV melting profiles,(Figure 2b; Tm= 53.4! 1.0 8C; n= 3); the opposite and unex-pected independent hybridization of the multicomponentDNA duplexes of different melting temperatures would haveled to a very broad transition. Furthermore, dynamic lightscattering (DLS) established the monomeric nature of thenanobarrels, as only a single peak with a hydrodynamic radiusof 5.5! 0.1 nm was observed (Figure 2c). The radius is largerthan the calculated value of 4.9 nm,[22,23] but in line with theaccuracy of DLS measurements for related DNA nano-structures.[23,24] The detailed dimensions of the nanobarrelwere established with atomic force microscopy (AFM)analysis (Figure 2d). The apparent height of (2.20!0.25) nm was expected for tip-compressed hollow DNAnanostructures.[8,25] Similarly, the AFM-derived length and

width of (20.4! 4.5) nm and (9.7! 2) nm (full width at halfmaximum), respectively, were in good agreement with thetheoretical dimensions (14 nm and 5.5 nm) after correctingfor tip-deconvolution.[10,26]

Having completed their structural characterization, weinvestigated whether the porphyrin-tagged DNA nanoporecan be stably anchored into lipid bilayers. In this examinationwe took advantage of the fluorescence properties of theporphyrin tag. Once incubated with giant unilamellar vesicles,the tagged nanopores could be visualized using fluorescencemicroscopy (Figure 3a, bottom). A bright-field image docu-mented the shape of the vesicle (Figure 3a, top). The brightfluorescence spots in the ring-shaped bilayer may representindividual DNA nanopores, given that the imaging conditionsof our custom-built set-up are powerful enough to detectsingle fluorophores.[27] However, as the DNA pores rapidlydiffuse out of the focal plane (Supporting Information,Movie SM1) and exhibit surprising photostability, it is hardto detect individual pores by single-fluorophore photobleach-ing. The anchoring of the nanopore into the bilayer wasconfirmed by independent spectroscopic analysis based on

Figure 1. DNA-nanopore-carrying porphyrin-based lipid anchors.a) Deoxyuridine bonded to tetraphenylporphyrin (TPP) through anacetylene linkage at the 5 position of the nucleobase. b) A DNAnanopore composed of six interconnected duplexes, represented ascylinders. The six-component DNA oligonucleotides are shown ingreen. The magenta porphyrin tags anchor the DNA nanopore into thelipid bilayer. For reasons of visual clarity, only the porphyrin corewithout the phenyl groups is shown. The TPP tags are not drawn toscale.

Figure 2. Characterization of porphyrin-modified nanobarrels assem-bled from DNA oligonucleotides. a) Agarose gel electrophoresis. Lane1, 100 bp marker; lane 2, DNA nanopore; lane 3, TPP-DNA nanopore.b) UV melting profile of a TPP nanopore. c) Dynamic light scatteringtrace of TPP nanopore. d) AFM micrograph of individual nanopores(arrows), and of elongated nanopore assemblies, which are stabilizedby inter-barrel base stacking; scale bar=50 nm. abs.=absorbance,T= temperature.

Figure 3. Porphyrin tags anchor DNA nanopores into lipid bilayers.a) Microscopic bright-field (top) and fluorescence (bottom;lexc=532 nm) images of a DPhPC vesicle containing DNA nanopores;scale bars=5 mm. b) The fluorescence emission spectrum for the TPPnanopore (c) is shifted upon insertion into vesicle bilayers (a ;lexc=424 nm). F= fluorescence intensity.

.AngewandteCommunications

12070 www.angewandte.org ! 2013 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim Angew. Chem. Int. Ed. 2013, 52, 12069 –12072

Angewandte Chemie 10.1002/anie.201305765

tional DNA, the hydrophobic belt features charge-masked alkyl-phosphorothioates (PPT) with an ethyl moiety attached to thethiol group (Figure 1A, inset), thereby annulling the typicalnegative charge of a phosphate anion. A total of 72 PPT groups(12 per each duplex) makes up the hydrophobic belt.The synthesis of the DNA barrel started with the chemical

modification of commercially available PPT-containing DNAoligonucleotides to generate charge-capped ethyl-PPT moieties(for sequences see Supporting Information, Table S-1). TheDNA strands were subjected to a modification protocol inwhich ethyl iodide reacts with the thiol group via nucleophilicsubstitution to yield the ethyl-protected PPT (Figure 2A).52

The chemical modification achieved a yield of 70% or 100% asdetermined by high-performance liquic chromatography(HPLC; Figure 2B and C), depending on the reactionconditions (see legend to Figure 2). The more challengingcomplete modification was also attained for DNA barrel strandswith 5 and 18 PPT groups (Supporting Information, Figure S-

4). Furthermore, the chemical change of the ethyl modificationwas successfully confirmed by mass spectrometry (SupportingInformation, Figure S-5). We note that, in the subsequentexperiments, barrels formed from 70% or 100% modified PPT-DNA strands had the same biophysical characteristics exceptthat the latter were of 5 °C lower thermal stability asdetermined by UV-melting profiles (Supporting Information,Figure S-6). In the following, we show data of the 70% barrels.The barrel was assembled by heating and cooling an

equimolar mixture of all component strands which includessix ethyl-PPT modified scaffold strands and eight staple strands(Figure 1B; Supporting Information, Table S-1 and Figure S-3).The purity and size of the assembly product were assessed bysize exclusion chromatography (SEC) and found to result in amajor chromatographic peak corresponding to a 70% yield offormation (Figure 3, blue line). In agreement with a hollowDNA barrel, the peak’s apparent MW of 500 kDaobtained bycomparison with molecularly compact protein standardsis

Figure 1. Schematic representation of a DNA nanopore composed of six interconnected duplexes represented as cylinders. (A) On the external face,the barrel features a membrane-spanning hydrophobic belt (magenta) where conventional phosphates of the DNA backbone are substituted bycharge-neutral phosphorothioate-ethyl groups (inset). For reasons of clarity, only one stereoisomer of the ethane-PPT group is drawn. (B) Map ofthe DNA nanostructure with six duplexes (numbered) being formed by six vertical scaffold strands (blue) and eight colored staple strands which runhorizontally in conventional honeycomb fashion. The stars indicate the position of the phosphorothioate groups.

Figure 2. Generation of ethyl-phosphorothioate DNA. (A) Reaction scheme for the modification of PPT-DNA with iodo-ethane. (B, C) HPLCtraces determine the extent of modification for a DNA strand carrying a single phosphorothioate group under reaction conditions (B) 10 equiv ofiodo-ethane, 55 °C, 1.5 h, and (C) 20 equiv of iodo-ethane, 65 °C, 1.5 h. Panel B shows the trace of the product mixture with unreacted (black) andethyl-modified phosphorothioate DNA (purple), while C displays the traces of starting material (black) and product mixture (purple). The elutiontimes of the peaks in B and C differ due to the use of two different elution gradients. The extent of modification was confirmed in two additionalindependent experiments (not shown).

Nano Letters Letter

dx.doi.org/10.1021/nl304147f | Nano Lett. 2013, 13, 2351−23562352

Nano Letters, 13: 2351 ~1 nS conductance

Langecker, M. et al., Science: 338, 932-936.

Page 35: Introduction to MD simulations of DNA systems › Training › Workshop › Urbana2018c › ... · 2019-02-05 · 1 2 4 6 10 2 4 6 100 ar) 22 24 26 28 30 32 DNA-DNA distance (Å)

�35

NanocapillaryNanocapillary

Göpfrich, Kerstin et al., Nano Lett., 2015, 15(5), 3134−3138.

A

MD simulation of DNA channel conductivityNanoEngineer-1caDNAno or

all-atomcoordinates

(Right) design from Burns, Jonathan R. et al., Nat. Nanotechnol., 2016, 11, 152−156.

Page 36: Introduction to MD simulations of DNA systems › Training › Workshop › Urbana2018c › ... · 2019-02-05 · 1 2 4 6 10 2 4 6 100 ar) 22 24 26 28 30 32 DNA-DNA distance (Å)

�36

NanocapillaryNanocapillary

Göpfrich, Kerstin et al., Nano Lett., 2015, 15(5), 3134−3138.

A

MD simulation of DNA channel conductivityDNA

Optimize

Cholesterol

Lipid

Mg(H2O)62+

H2O

K+

Cl-

Page 37: Introduction to MD simulations of DNA systems › Training › Workshop › Urbana2018c › ... · 2019-02-05 · 1 2 4 6 10 2 4 6 100 ar) 22 24 26 28 30 32 DNA-DNA distance (Å)

�37

NanocapillaryNanocapillary

Göpfrich, Kerstin et al., Nano Lett., 2015, 15(5), 3134−3138.

A

Instantaneous current:

MD simulation of DNA channel conductivityDNA

Optimize

Cholesterol

Lipid

Mg(H2O)62+

H2O

K+

Cl-

E

Page 38: Introduction to MD simulations of DNA systems › Training › Workshop › Urbana2018c › ... · 2019-02-05 · 1 2 4 6 10 2 4 6 100 ar) 22 24 26 28 30 32 DNA-DNA distance (Å)

�38

NanocapillaryNanocapillary

Göpfrich, Kerstin et al., Nano Lett., 2015, 15(5), 3134−3138.

A

MD simulation of DNA channel conductivity

E

Page 39: Introduction to MD simulations of DNA systems › Training › Workshop › Urbana2018c › ... · 2019-02-05 · 1 2 4 6 10 2 4 6 100 ar) 22 24 26 28 30 32 DNA-DNA distance (Å)

Chemistry matters

tional DNA, the hydrophobic belt features charge-masked alkyl-phosphorothioates (PPT) with an ethyl moiety attached to thethiol group (Figure 1A, inset), thereby annulling the typicalnegative charge of a phosphate anion. A total of 72 PPT groups(12 per each duplex) makes up the hydrophobic belt.The synthesis of the DNA barrel started with the chemical

modification of commercially available PPT-containing DNAoligonucleotides to generate charge-capped ethyl-PPT moieties(for sequences see Supporting Information, Table S-1). TheDNA strands were subjected to a modification protocol inwhich ethyl iodide reacts with the thiol group via nucleophilicsubstitution to yield the ethyl-protected PPT (Figure 2A).52

The chemical modification achieved a yield of 70% or 100% asdetermined by high-performance liquic chromatography(HPLC; Figure 2B and C), depending on the reactionconditions (see legend to Figure 2). The more challengingcomplete modification was also attained for DNA barrel strandswith 5 and 18 PPT groups (Supporting Information, Figure S-

4). Furthermore, the chemical change of the ethyl modificationwas successfully confirmed by mass spectrometry (SupportingInformation, Figure S-5). We note that, in the subsequentexperiments, barrels formed from 70% or 100% modified PPT-DNA strands had the same biophysical characteristics exceptthat the latter were of 5 °C lower thermal stability asdetermined by UV-melting profiles (Supporting Information,Figure S-6). In the following, we show data of the 70% barrels.The barrel was assembled by heating and cooling an

equimolar mixture of all component strands which includessix ethyl-PPT modified scaffold strands and eight staple strands(Figure 1B; Supporting Information, Table S-1 and Figure S-3).The purity and size of the assembly product were assessed bysize exclusion chromatography (SEC) and found to result in amajor chromatographic peak corresponding to a 70% yield offormation (Figure 3, blue line). In agreement with a hollowDNA barrel, the peak’s apparent MW of 500 kDaobtained bycomparison with molecularly compact protein standardsis

Figure 1. Schematic representation of a DNA nanopore composed of six interconnected duplexes represented as cylinders. (A) On the external face,the barrel features a membrane-spanning hydrophobic belt (magenta) where conventional phosphates of the DNA backbone are substituted bycharge-neutral phosphorothioate-ethyl groups (inset). For reasons of clarity, only one stereoisomer of the ethane-PPT group is drawn. (B) Map ofthe DNA nanostructure with six duplexes (numbered) being formed by six vertical scaffold strands (blue) and eight colored staple strands which runhorizontally in conventional honeycomb fashion. The stars indicate the position of the phosphorothioate groups.

Figure 2. Generation of ethyl-phosphorothioate DNA. (A) Reaction scheme for the modification of PPT-DNA with iodo-ethane. (B, C) HPLCtraces determine the extent of modification for a DNA strand carrying a single phosphorothioate group under reaction conditions (B) 10 equiv ofiodo-ethane, 55 °C, 1.5 h, and (C) 20 equiv of iodo-ethane, 65 °C, 1.5 h. Panel B shows the trace of the product mixture with unreacted (black) andethyl-modified phosphorothioate DNA (purple), while C displays the traces of starting material (black) and product mixture (purple). The elutiontimes of the peaks in B and C differ due to the use of two different elution gradients. The extent of modification was confirmed in two additionalindependent experiments (not shown).

Nano Letters Letter

dx.doi.org/10.1021/nl304147f | Nano Lett. 2013, 13, 2351−23562352

previously published nanopore,[10] but the current design isconsiderably simpler, as six rather than 14 strands are used.Crucially, two porphyrin tags are positioned at the end of thenanobarrel to achieve its insertion into the bilayer in a direc-tional manner (Figure 1b). To facilitate the comparison to ourprevious nanopore work, we used a design with two DNAstrands that contain phosphorothiate groups in place of thephosphates in a number of positions in the backbone. Thethioate groups behave like the native negatively chargedphosphate groups at our experimental conditions of pH 8.0(Table S1).[10]

The DNA nanopore was assembled by heating andcooling an equimolar mixture of four regular and two TPP-modified DNA strands (for sequences see Table S1). Theassembly mixture was characterized to confirm the correctand successful formation of the DNA nanopore. Native gelelectrophoresis yielded a band, which migrated to the sameheight as a control nanopore without the porphyrin anchor(Figure 2a, lanes 2 and 3, respectively; main band co-migrat-ing at the 550 bp marker). The tailing of the band for theporphyrin DNA pore does not indicate unfolding, but israther caused by the very hydrophobic tag.[9,19,20] Theconcerted assembly into the nanobarrels was also confirmedby a single defined transition in the UV melting profiles,(Figure 2b; Tm= 53.4! 1.0 8C; n= 3); the opposite and unex-pected independent hybridization of the multicomponentDNA duplexes of different melting temperatures would haveled to a very broad transition. Furthermore, dynamic lightscattering (DLS) established the monomeric nature of thenanobarrels, as only a single peak with a hydrodynamic radiusof 5.5! 0.1 nm was observed (Figure 2c). The radius is largerthan the calculated value of 4.9 nm,[22,23] but in line with theaccuracy of DLS measurements for related DNA nano-structures.[23,24] The detailed dimensions of the nanobarrelwere established with atomic force microscopy (AFM)analysis (Figure 2d). The apparent height of (2.20!0.25) nm was expected for tip-compressed hollow DNAnanostructures.[8,25] Similarly, the AFM-derived length and

width of (20.4! 4.5) nm and (9.7! 2) nm (full width at halfmaximum), respectively, were in good agreement with thetheoretical dimensions (14 nm and 5.5 nm) after correctingfor tip-deconvolution.[10,26]

Having completed their structural characterization, weinvestigated whether the porphyrin-tagged DNA nanoporecan be stably anchored into lipid bilayers. In this examinationwe took advantage of the fluorescence properties of theporphyrin tag. Once incubated with giant unilamellar vesicles,the tagged nanopores could be visualized using fluorescencemicroscopy (Figure 3a, bottom). A bright-field image docu-mented the shape of the vesicle (Figure 3a, top). The brightfluorescence spots in the ring-shaped bilayer may representindividual DNA nanopores, given that the imaging conditionsof our custom-built set-up are powerful enough to detectsingle fluorophores.[27] However, as the DNA pores rapidlydiffuse out of the focal plane (Supporting Information,Movie SM1) and exhibit surprising photostability, it is hardto detect individual pores by single-fluorophore photobleach-ing. The anchoring of the nanopore into the bilayer wasconfirmed by independent spectroscopic analysis based on

Figure 1. DNA-nanopore-carrying porphyrin-based lipid anchors.a) Deoxyuridine bonded to tetraphenylporphyrin (TPP) through anacetylene linkage at the 5 position of the nucleobase. b) A DNAnanopore composed of six interconnected duplexes, represented ascylinders. The six-component DNA oligonucleotides are shown ingreen. The magenta porphyrin tags anchor the DNA nanopore into thelipid bilayer. For reasons of visual clarity, only the porphyrin corewithout the phenyl groups is shown. The TPP tags are not drawn toscale.

Figure 2. Characterization of porphyrin-modified nanobarrels assem-bled from DNA oligonucleotides. a) Agarose gel electrophoresis. Lane1, 100 bp marker; lane 2, DNA nanopore; lane 3, TPP-DNA nanopore.b) UV melting profile of a TPP nanopore. c) Dynamic light scatteringtrace of TPP nanopore. d) AFM micrograph of individual nanopores(arrows), and of elongated nanopore assemblies, which are stabilizedby inter-barrel base stacking; scale bar=50 nm. abs.=absorbance,T= temperature.

Figure 3. Porphyrin tags anchor DNA nanopores into lipid bilayers.a) Microscopic bright-field (top) and fluorescence (bottom;lexc=532 nm) images of a DPhPC vesicle containing DNA nanopores;scale bars=5 mm. b) The fluorescence emission spectrum for the TPPnanopore (c) is shifted upon insertion into vesicle bilayers (a ;lexc=424 nm). F= fluorescence intensity.

.AngewandteCommunications

12070 www.angewandte.org ! 2013 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim Angew. Chem. Int. Ed. 2013, 52, 12069 –12072

previously published nanopore,[10] but the current design isconsiderably simpler, as six rather than 14 strands are used.Crucially, two porphyrin tags are positioned at the end of thenanobarrel to achieve its insertion into the bilayer in a direc-tional manner (Figure 1b). To facilitate the comparison to ourprevious nanopore work, we used a design with two DNAstrands that contain phosphorothiate groups in place of thephosphates in a number of positions in the backbone. Thethioate groups behave like the native negatively chargedphosphate groups at our experimental conditions of pH 8.0(Table S1).[10]

The DNA nanopore was assembled by heating andcooling an equimolar mixture of four regular and two TPP-modified DNA strands (for sequences see Table S1). Theassembly mixture was characterized to confirm the correctand successful formation of the DNA nanopore. Native gelelectrophoresis yielded a band, which migrated to the sameheight as a control nanopore without the porphyrin anchor(Figure 2a, lanes 2 and 3, respectively; main band co-migrat-ing at the 550 bp marker). The tailing of the band for theporphyrin DNA pore does not indicate unfolding, but israther caused by the very hydrophobic tag.[9,19,20] Theconcerted assembly into the nanobarrels was also confirmedby a single defined transition in the UV melting profiles,(Figure 2b; Tm= 53.4! 1.0 8C; n= 3); the opposite and unex-pected independent hybridization of the multicomponentDNA duplexes of different melting temperatures would haveled to a very broad transition. Furthermore, dynamic lightscattering (DLS) established the monomeric nature of thenanobarrels, as only a single peak with a hydrodynamic radiusof 5.5! 0.1 nm was observed (Figure 2c). The radius is largerthan the calculated value of 4.9 nm,[22,23] but in line with theaccuracy of DLS measurements for related DNA nano-structures.[23,24] The detailed dimensions of the nanobarrelwere established with atomic force microscopy (AFM)analysis (Figure 2d). The apparent height of (2.20!0.25) nm was expected for tip-compressed hollow DNAnanostructures.[8,25] Similarly, the AFM-derived length and

width of (20.4! 4.5) nm and (9.7! 2) nm (full width at halfmaximum), respectively, were in good agreement with thetheoretical dimensions (14 nm and 5.5 nm) after correctingfor tip-deconvolution.[10,26]

Having completed their structural characterization, weinvestigated whether the porphyrin-tagged DNA nanoporecan be stably anchored into lipid bilayers. In this examinationwe took advantage of the fluorescence properties of theporphyrin tag. Once incubated with giant unilamellar vesicles,the tagged nanopores could be visualized using fluorescencemicroscopy (Figure 3a, bottom). A bright-field image docu-mented the shape of the vesicle (Figure 3a, top). The brightfluorescence spots in the ring-shaped bilayer may representindividual DNA nanopores, given that the imaging conditionsof our custom-built set-up are powerful enough to detectsingle fluorophores.[27] However, as the DNA pores rapidlydiffuse out of the focal plane (Supporting Information,Movie SM1) and exhibit surprising photostability, it is hardto detect individual pores by single-fluorophore photobleach-ing. The anchoring of the nanopore into the bilayer wasconfirmed by independent spectroscopic analysis based on

Figure 1. DNA-nanopore-carrying porphyrin-based lipid anchors.a) Deoxyuridine bonded to tetraphenylporphyrin (TPP) through anacetylene linkage at the 5 position of the nucleobase. b) A DNAnanopore composed of six interconnected duplexes, represented ascylinders. The six-component DNA oligonucleotides are shown ingreen. The magenta porphyrin tags anchor the DNA nanopore into thelipid bilayer. For reasons of visual clarity, only the porphyrin corewithout the phenyl groups is shown. The TPP tags are not drawn toscale.

Figure 2. Characterization of porphyrin-modified nanobarrels assem-bled from DNA oligonucleotides. a) Agarose gel electrophoresis. Lane1, 100 bp marker; lane 2, DNA nanopore; lane 3, TPP-DNA nanopore.b) UV melting profile of a TPP nanopore. c) Dynamic light scatteringtrace of TPP nanopore. d) AFM micrograph of individual nanopores(arrows), and of elongated nanopore assemblies, which are stabilizedby inter-barrel base stacking; scale bar=50 nm. abs.=absorbance,T= temperature.

Figure 3. Porphyrin tags anchor DNA nanopores into lipid bilayers.a) Microscopic bright-field (top) and fluorescence (bottom;lexc=532 nm) images of a DPhPC vesicle containing DNA nanopores;scale bars=5 mm. b) The fluorescence emission spectrum for the TPPnanopore (c) is shifted upon insertion into vesicle bilayers (a ;lexc=424 nm). F= fluorescence intensity.

.AngewandteCommunications

12070 www.angewandte.org ! 2013 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim Angew. Chem. Int. Ed. 2013, 52, 12069 –12072

Angewandte Chemie 10.1002/anie.201305765

Nano Letters, 13: 2351

= K+ = Cl– E

Cholesterol

0 ns26 ns89 ns162 ns

RMSF < 3 Å

RMSF > 6 Å

Density of lipid head groups

Page 40: Introduction to MD simulations of DNA systems › Training › Workshop › Urbana2018c › ... · 2019-02-05 · 1 2 4 6 10 2 4 6 100 ar) 22 24 26 28 30 32 DNA-DNA distance (Å)

Small conductance DNA channel

Gramicidinion channels

Conductance < 0.1 nS

Y

Goepfrich, et al., Nano Lett 16: 4665 (2016)

Figure 4: Molecular dynamics simulations of the DNA funnel and the duplex. A) All-atommodels of the funnel (top, 7,963,516 atoms) and the duplex (bottom, 140,630 atoms) with hy-drophobic tags (red) embedded in a lipid membrane (green) before equilibration. Each systemcontains magnesium ions in amount sufficient to neutralize the electrical charge of the DNAchannel and 1M KCl solution (not shown). Electric field (direction of positive transmembranebias) as indicated. B) Steady-state local densities of lipid chain (carbon atoms, green), DNA(phosphorus atoms, blue) and ion current (streamlines). The arrows indicate the direction ofthe local ionic current flux, the color the flux’ magnitude. The units of the blue, green andpurple-red-yellow color bars are molar (M), molar (M) and nA/nm2, respectively. The mapswere computed from MD trajectories of the DNA funnel (top, 19.2 ns) and the DNA duplex(bottom, 672 ns) at a +100 mV bias sampled with a frequency of 240 ps per frame; radiallyaveraged about the z-axis to improve the resolution. C) All-point conductance histograms withlogarithmic binning for both constructs. The conductance histograms were computed usingblock averaged instantaneous currents at +100, +30, �30, and �100 mV transmembrane bi-ases; the block-average size was 2.88 ns (top) and 9.6 (bottom) ns. The dashed lines indicatesthe mean conductance value. D) Current-voltage characteristics of the funnel (top) and the du-plex (bottom). The data is block averaged with a block size of 2.88 ns (top) and 9.6 ns (bottom).Error bars indicate the standard error of the mean, the dashed line represents a linear fit.

8

Simulation

Pro

babi

lity

Figure 3: Ionic current recordings of the cholesterol-tagged funnel (top) and the porphyrin-tagged duplex (bottom) in 1M KCl, 10mM MES, pH 6.0. A) Exemplary current traces showingtwo consecutive insertions of the funnel (top, recorded at 10mV) and closures of the duplex(bottom, recorded at 50mV). B) Histogram of insertion and closure steps for the funnel andthe duplex, logarithmic binning was used for both. Dashed lines at 30 nS for the funnel and0.13 nS for the duplex represent the means of the histograms C) Current-voltage characteristicsof stable insertions of the funnel (n = 6) and the duplex (n = 8). Error bars correspond to thestandard deviation of n independent recordings. The dashed line represents a linear fit.

insertions can be stable for hours, we also observe transient insertions or insertion-attempts,

all contributing to the heterogeneity of observed conductances, fig. S8-S10. The IV-traces in

Fig. 3C combine measurements where the pores did not switch conductance states during the

recording. The dashed lines represent a linear fit with a gradient of GIVf = 20 nS for the funnel

and GIVd = 0.1 nS for the duplex.

Our observations with the DNA duplex settle the ongoing debate regarding the pathway of ions

through DNA membrane pores (5,8,21). Studies of a genetically engineered ↵-hemolysin pore

with a truncated stem (22) as well as reports of protein-induced lipid ion channels (23) suggest

that alternative conductance pathways without a physical channel may also exist in biological

systems.

6

Experiment

Cou

nt

Figure 4: Molecular dynamics simulations of the DNA funnel and the duplex. A) All-atommodels of the funnel (top, 7,963,516 atoms) and the duplex (bottom, 140,630 atoms) with hy-drophobic tags (red) embedded in a lipid membrane (green) before equilibration. Each systemcontains magnesium ions in amount sufficient to neutralize the electrical charge of the DNAchannel and 1M KCl solution (not shown). Electric field (direction of positive transmembranebias) as indicated. B) Steady-state local densities of lipid chain (carbon atoms, green), DNA(phosphorus atoms, blue) and ion current (streamlines). The arrows indicate the direction ofthe local ionic current flux, the color the flux’ magnitude. The units of the blue, green andpurple-red-yellow color bars are molar (M), molar (M) and nA/nm2, respectively. The mapswere computed from MD trajectories of the DNA funnel (top, 19.2 ns) and the DNA duplex(bottom, 672 ns) at a +100 mV bias sampled with a frequency of 240 ps per frame; radiallyaveraged about the z-axis to improve the resolution. C) All-point conductance histograms withlogarithmic binning for both constructs. The conductance histograms were computed usingblock averaged instantaneous currents at +100, +30, �30, and �100 mV transmembrane bi-ases; the block-average size was 2.88 ns (top) and 9.6 (bottom) ns. The dashed lines indicatesthe mean conductance value. D) Current-voltage characteristics of the funnel (top) and the du-plex (bottom). The data is block averaged with a block size of 2.88 ns (top) and 9.6 ns (bottom).Error bars indicate the standard error of the mean, the dashed line represents a linear fit.

8

Experiment: Ulrich Keyser (Cambridge) Eugen Stulz (Southampton) Mathias Winterhalter (Jacobs U)

Conductance: ~ 0.1 nS

Cur

rent

/ nA

Cur

rent

/ nA

1 M KCl

1 helix 140,000 atoms

Page 41: Introduction to MD simulations of DNA systems › Training › Workshop › Urbana2018c › ... · 2019-02-05 · 1 2 4 6 10 2 4 6 100 ar) 22 24 26 28 30 32 DNA-DNA distance (Å)

All-atom MD simulation of lipid-DNA interface

Lipid molecules can translocate to the other leaflet through the toroidal pore made by DNA

Dis

tanc

e fr

om p

ore

Z

No applied electric

field

Distance from pore

Chen Yu Li

Nature Communications 9: 2426 (2018)

Page 42: Introduction to MD simulations of DNA systems › Training › Workshop › Urbana2018c › ... · 2019-02-05 · 1 2 4 6 10 2 4 6 100 ar) 22 24 26 28 30 32 DNA-DNA distance (Å)

Lipid translocation through toroidal pores is very common and very fast

Page 43: Introduction to MD simulations of DNA systems › Training › Workshop › Urbana2018c › ... · 2019-02-05 · 1 2 4 6 10 2 4 6 100 ar) 22 24 26 28 30 32 DNA-DNA distance (Å)

�43

Lipid translocation in cells is catalyzed by enzyme

PC Sph PE PS

Extracellular space

CytoplasmImage credit:

Kelly Yang

Lipid molecules are asymmetrically distributed in the cell membrane.

apoptosis or thrombin formation:

scramblaseDeficiency in lipid scrambling could result in autoimmune

response or Scott syndrome.

F(t

)/F

(0)

Time (t)

1

0

Time (t)

F(t

)/F

(0)

1

0

0.5

xin: ratio of labeled lipid in the inner leafletkq and ks: rate constant for quenching and scrambling

Brunner et al, Nature 516, 207–212, 2014

Page 44: Introduction to MD simulations of DNA systems › Training › Workshop › Urbana2018c › ... · 2019-02-05 · 1 2 4 6 10 2 4 6 100 ar) 22 24 26 28 30 32 DNA-DNA distance (Å)

Experimental verificationAlex Ohmann

Ohmann, Li, … Ulrich F. Keyser, Aksimentiev, Nature Communications 9: 2426 (2018)

Page 45: Introduction to MD simulations of DNA systems › Training › Workshop › Urbana2018c › ... · 2019-02-05 · 1 2 4 6 10 2 4 6 100 ar) 22 24 26 28 30 32 DNA-DNA distance (Å)

Experimental verification

Ohmann, Li, … Ulrich F. Keyser, Aksimentiev, Nature Communications 9: 2426 (2018)

Scale bars are 10 µm Alex Ohmann

Page 46: Introduction to MD simulations of DNA systems › Training › Workshop › Urbana2018c › ... · 2019-02-05 · 1 2 4 6 10 2 4 6 100 ar) 22 24 26 28 30 32 DNA-DNA distance (Å)

Works in human cellsHuman cells contain PS lipids at the inner membraneAnnexin V binds specifically to PS lipids

Breast cancer cells from the cell line MDA-MB-231 Positive control: apoptosis-inducing microbial alkaloid staurosporineNegative control: DNA folding buffer

Scale bar is 20 µm

Nature Communications 9: 2426 (2018)

Page 47: Introduction to MD simulations of DNA systems › Training › Workshop › Urbana2018c › ... · 2019-02-05 · 1 2 4 6 10 2 4 6 100 ar) 22 24 26 28 30 32 DNA-DNA distance (Å)

Tutorials overviewDesign Simulations

Cadnano (today)

Cadnano toolkit (tomorrow)

All-atom with NAMD (today)

Coarse-grained with ARBD (tomorrow)

Images and movies with VMD

Page 48: Introduction to MD simulations of DNA systems › Training › Workshop › Urbana2018c › ... · 2019-02-05 · 1 2 4 6 10 2 4 6 100 ar) 22 24 26 28 30 32 DNA-DNA distance (Å)

Acknowledgements

JejoongYooChen-YuLiJejoong Yoo

Chen-Yu Li