• No results found

Capping Ligand Vortices as “Atomic Orbitals” in Nanocrystal Self-Assembly

N/A
N/A
Protected

Academic year: 2020

Share "Capping Ligand Vortices as “Atomic Orbitals” in Nanocrystal Self-Assembly"

Copied!
31
0
0

Loading.... (view fulltext now)

Full text

(1)

Ames Laboratory Accepted Manuscripts Ames Laboratory

10-27-2017

Capping Ligand Vortices as “Atomic Orbitals” in

Nanocrystal Self-Assembly

Curt Waltmann

Iowa State University and Ames Laboratory, waltmann@iastate.edu

Nathan Horst

Iowa State University and Ames Laboratory, nhorst@iastate.edu

Alex Travesset

Iowa State University and Ames Laboratory, trvsst@ameslab.gov

Follow this and additional works at:http://lib.dr.iastate.edu/ameslab_manuscripts

Part of theEngineering Physics Commons,Materials Science and Engineering Commons, and theNanoscience and Nanotechnology Commons

This Article is brought to you for free and open access by the Ames Laboratory at Iowa State University Digital Repository. It has been accepted for inclusion in Ames Laboratory Accepted Manuscripts by an authorized administrator of Iowa State University Digital Repository. For more information, please contactdigirep@iastate.edu.

Recommended Citation

Waltmann, Curt; Horst, Nathan; and Travesset, Alex, "Capping Ligand Vortices as “Atomic Orbitals” in Nanocrystal Self-Assembly" (2017).Ames Laboratory Accepted Manuscripts. 94.

(2)

Capping Ligand Vortices as “Atomic Orbitals” in Nanocrystal

Self-Assembly

Abstract

We present a detailed analysis of the interaction between two nanocrystals capped with ligands consisting of hydrocarbon chains by united atom molecular dynamics simulations. We show that the bonding of two nanocrystals is characterized by ligand textures in the form of vortices. These results are generalized to nanocrystals of different types (differing core and ligand sizes) where the structure of the vortices depends on the softness asymmetry. We provide rigorous calculations for the binding free energy, show that these energies are independent of the chemical composition of the cores, and derive analytical formulas for the equilibrium separation. We discuss the implications of our results for the self-assembly of single-component and binary nanoparticle superlattices. Overall, our results show that the structure of the ligands completely determines the bonding of nanocrystals, fully supporting the predictions of the recently proposed Orbifold topological model.

Keywords

nanocrystal, self-assembly, skyrmions, superlattices, thermodynamics, vortices

Disciplines

(3)

Capping Ligand Vortices as “Atomic Orbitals” in

Nanocrystal Self-Assembly

Curt Waltmann,

Nathan Horst,

†,¶

and Alex Travesset

∗,‡,†

†Department of Materials Science and Engineering and Ames lab, Ames, IA, 50011, USA

‡Department of Physics and Astronomy and Ames lab, Ames, IA 50011, USA

¶Also a first author

E-mail: trvsst@ameslab.gov

Abstract

We present a detailed analysis of the interaction between two nanocrystals capped

with ligands consisting of hydrocarbon chains by united atom molecular dynamics

simulations. We show that the bonding of two nanocrystals is characterized by

lig-and textures in the form of vortices. These results are generalized to nanocrystals

of different types (differing core and ligand sizes) where the structure of the vortices

depends on the softness asymmetry. We provide rigorous calculations for the binding

free energy, show that these energies are independent of the chemical composition of

the cores and derive analytical formulas for the equilibrium separation. We discuss

the implications of our results for the self-assembly of single component and binary

nanoparticle superlattices. Overall, our results show that the structure of the ligands

completely determines the bonding of nanocrystals, fully supporting the predictions of

(4)

Keywords

self-assembly, nanocrystal, thermodynamics, superlattices, skyrmions, vortices

Nanocrystals (NCs) are nanometer scale crystalline inorganic cores that have been

func-tionalized by capping ligands. Assembling these NCs into materials that perform new and

potentially revolutionary functions is one of the central goals in nanotechnology. The

cap-ping ligand is necessary to direct and program the assembly of NCs into relevant

struc-tures, and plays a role akin to electronic orbitals in materials made of simple elements and

molecules. Typical ligands include DNA, polymers or hydrocarbons,1 leading to periodic and

quasi-periodic structures of NCs,i.e. superlattices, self-assembled by DNA-hybridization,2,3

[image:4.612.74.534.331.653.2]

electrostatic phase separation4–6 or solvent evaporation,7–18 just to name a few.

Figure 1: NCs as soft skyrmions: Hedgehog and vortex, with the different parameters Ai

(5)

Superlattices involving two NC species, i.e. binary nanoparticle superlattices (BNSLs),

are among the most sophisticated structures assembled to date, and more than twenty

dif-ferent structures of BNSLs in 2D and 3D have already been reported.19 The first theoretical

description of these BNSLs followed previous approaches for micron-size colloids,20 where

NCs are modeled as hard spheres (HS) whose diameter is given by the NC nearest

neigh-bor distance of a single component two dimensional triangular lattice. From these studies,

a correlation between maxima of the HS packing fraction and the presumably equilibrium

structures was obtained.12,21 Yet, such correlation is somewhat rudimentary, as many BNSLs

exist at low HS packing fraction and, furthermore, lattice distances are, more often than

not, inconsistent with the assumed HS radius.22 Attempts to describe those disagreements

in terms of more flexible potentials23,24 or other coarse-grained models25 expanded the

pre-dictable range of equilibrium lattices, but did not resolve the flaws of the HS model.22 It

is only with the advent of the Orbifold Topological Model (OTM),22 that a full and

rigor-ous description of the structural properties of BNSLs was achieved, as demonstrated by the

successful parameter free description of more than 160 experiments.26

The OTM describes each NC as a “Soft Skyrmion”, see Fig. 1. IfLdenotes the maximum

extension of the capping ligand, Ri the core radius and ri the NC radius, the dimensionless

variables

λi =

L Ri

, τOP M =

ri

Ri

, (1)

are introduced. The actual NC radius is given by the Optimal Packing Model(OPM)

for-mula27

τOP M = (1 + 3ξiλi)1/3 , (2)

where ξi = AA0

i,A0 is the smallest possible molecular area and Ai the actual molecular area.

The OPM value will be referred to as the NC radius ri and is generally smaller than the

hydrodynamic radiusrH

i defined as the radius of the free (unperturbed) NC, with associated

variable τH

(6)

OTM predicts a smaller radius ¯ri < ri, enabled by vortex textures, see Fig. 1. A detailed

description of how to calculate ¯ri is provided in the original references of the OTM.22,26

For two isolated NCs, the equilibrium distance ˜ri corresponds to the minimum of the

potential of Mean Force (PMF), and has been proposed to be described by the Overlap

Cone Model (OCM)28

τOCM =

1 +λi

2 −1 +

s

1 + 8(1 + 3ξiλi) (1 +λi)3

!

, (3)

where τOCM = ˜ri/Ri < τOP M, see Figure. 1. Since the case of an isolated pair of NCs

corresponds to a coordination of one, the OTM predicts that the NC “atomic orbitals”

created by deformed ligand structures can be realized as two vortex textures, one on each

NC, contained within the OCM cone, see Fig. 1.

In this paper, we analyze the interaction between two NCs within a united atom model

and analyze the postulated vortex configurations, thus identifying further the analogy of

cap-ping ligands as “atomic orbitals”. We include the case of two NCs with different radii and

ex-plore the long hydrocarbon chains investigated in real experiments. Previous studies focused

primarily on the potential of mean force, starting with those performed by Schapotschnikow

et al.28–30 Subsequent investigations have studied the role of solvent,31,32 more detailed

den-sity functional theories,33 as well as different approaches to the PMF34 and three body

effects.35 The relation of this work to our results is discussed further below.

Results

Model Parameters

We consider NCs whose crystalline cores consist of gold atoms and whose ligands are

(7)

eliminate hydrogens from our naming convention, instead denoting each one of the CH2 or

CH3 united atoms in our model as C. We studied three core sizes Au140(SCj)62, Au201(SCj)80

and Au1289(SCj)258, considering three distinct chain lengths of j = 9, 12 and 19 carbons. The

shapes of the cores follow truncated octahedral or cuboctahedral geometries, resulting in the

[111] and [001] families of planes as the exposed surfaces of the fcc crystalline cores (Figure

[image:7.612.132.466.222.377.2]

2).

Figure 2: Nanocrystal cores Au1289(SCj)258 (left), Au201(SCj)80 (bottom right), and

Au140(SCj)62 (top right) with Au atoms shown in yellow and S atoms shown in red.

Hydro-carbon chains are not drawn. The Au-Au distance is ∼ 2.8 ˚A. The geometry of the cores expose [001] and [111] planes. Average S-S distances are ∼ 4.1 ˚A. HC chains are attached at the sulfur sites, described by maximum length Li and having n =j −2 defined dihedral

angles where j is the number of carbons in the chain.

We follow the work of Luedtkeet al.36and choose sulfur binding sites as geometric points

on the surface of the core. These binding sites were grouped into three types: hollow, bridge,

and on-top, as described in the work of Tachibana et al.,37 based on the relative location of

the site to the close packed spheres of Au that comprise the core surface. The final positions

of the bound sulfur atoms were determined by adding a distance normal to the plane of the

core surface, where the distance is dependent on the type of binding site.37 The resulting

sulfur binding configuration on each nanocrystal core is shown in Figure 2. We note that

all ligand grafting densities studied in this work are high, close to the maximum allowed, so

grafting density will not be one of the parameters varied in this study.

(8)

assumptions. We define it by 2Ri =Di =Di,s+dAu−S whereDi,s is the diameter of a sphere

with equal surface area to that of the core, and dAu−S is the distance between the sulfur

atoms to the surface normal of the gold face. Another reasonable definition D0

i is given as

twice the radius of gyration of the sulfur atoms, with an additional term of 0.24˚A added due

to relative lengths of the S-C and C-C bonds. We define grafting densities as per molecule

areas of grafting sites, and note that the difference between our two definitions of the core

[image:8.612.69.544.267.374.2]

is negligible for all practical purposes, as shown in Table 1.

Table 1: Core Diameters and Grafting Densities, see the main text for description.

2Ri =Di(˚A) 2R0i =D

0

i(˚A) σ(˚A

−2) σ’(˚A−2)

Au140(SCm)62 18.57 18.33 17.47 17.02

Au201(SCm)80 20.30 20.22 16.18 16.06

Au1289(SCm)258 40.60 40.44 20.07 19.91

We use a united-atom model for the hydrocarbon surfactant chains, with each sulfur

and carbon in the chain grouped with their surrounding bonded hydrogen atoms into one

effective “atom”. Our force field follows Hautman and Klein,38 see SI for further details.

Table 2: NC Parameters Ri, L, λi. Hydrodynamic radius ri,mH and minimum of the PMF

2˜ri,m, the subscript m is meant to remind that this is the result found in the simulation.

Also, the same radii are given in the variableτ.

NC Ri(˚A) L λi rHi,m(˚A) 2˜ri,m τi,mH τ˜i,m

Au140(SC9)62 9.285 13.42 1.445 19.2 23.8 2.07 1.28

Au140(SC12)62 9.285 17.22 1.855 20.9 23.7 2.25 1.27

Au140(SC19)62 9.285 26.10 2.811 23.7 24.5 2.55 1.32

Au201(SC9)80 10.15 13.42 1.322 20.5 27.7 2.02 1.36

Au201(SC12)80 10.15 17.22 1.697 22.3 27.1 2.20 1.33

Au201(SC19)80 10.15 26.10 2.571 27.4 27.6 2.70 1.36

Au1289(SC9)258 20.30 13.42 0.661 29.5 49.0 1.45 1.21

Au1289(SC12)258 20.30 17.22 0.848 31.8 49.5 1.57 1.22

[image:8.612.71.545.534.684.2]
(9)

Configurational Characteristics

Potential of Mean Force

We only discuss the PMF in vacuum, i.e. for dry conditions, in the absence of solvent, and

included the nine cases described in Table 2. Note that this considers a range of λi, see

Eq. 1, over the interval [0.66,2.81]. Four representative cases for the PMF calculation are

provided in Fig. 3, see SI for the remaining cases. The method followed to perform these

calculations is described briefly in Materials and Methods and in more detail within the SI.

Our PMF figures clearly illustrate the competition between the attractive Van der Waals

forces and the entropically driven chain repulsion. The binding free energies are very strong,

in the range of several hundreds of kBT, an issue that we further discuss below. Note that

even at distances ρ ≈ 2rH

i,m there is a very significant attraction between NCs, driven by

the stretching of the chains towards each other. This effect becomes quite obvious for the

Au1289(SC19)258 NC, as shown in Fig. 3. Note that we introduce the subscript ”m” to our

notation to denote results calculated directly from our simulations.

Chain Structure

The OCM angle is given by:

cos(χOCM) =

˜ ri

Ri+Li

= τOCM

1 +λi

, (4)

and defines a cone where the hydrocarbons fill the entire space (melt condition), drawn as

a solid red cone in Fig. 1. This same cone is illustrated in Fig. 4 as a purple conical region.

A plot of the percentage of dihedrals in the trans configuration within the OCM cone is

provided in Fig. 4, where a noticeable decrease in the percentage of dihedrals in the trans

state occurs for the first and the third dihedrals in the inner OCM cone, χinner =χOCM/2,

depicted in cyan, where the numbering of dihedrals is elucidated in Fig. 2. This decreasing

(10)
[image:10.612.80.528.194.541.2]
(11)

main conformational changes are associated with the chains next to the center of the cone.

In Fig. 5 we present averaged configurations for both free and interacting NCs (taken at

the minimum of the free energy), shown from the side and from the front. For the case of a NC

in free space, the configuration resembles a hedgehog, where all the chains lie perpendicular

to the core (neutral lines or surfaces), while for the interacting case, they consist of chains

radiating from a central point with a non-zero projection over the core, denoted as vortices,

see Fig. 1. Note that beyond the OCM-cone, the configurations of the chain are very much

the same as for the free NC, resembling hedgehog configurations or neutral lines. It is also

apparent that there is negligible chain interdigitation among chains belonging to different

[image:11.612.94.517.315.579.2]

NCs, with chains entirely contained to the right or the left of the interaction plane.

Figure 4: Representation of the OCM and the inner OCM cone. The breaking of φ1 and φ3

(12)
[image:12.612.81.537.136.292.2]

Figure 5: Average configuration for the Au140(SC9)62 (a) average free particle, consisting of

a neutral surface (b) side view of the vortex at pair potential minimum (c) view of NC at the minimum of the pair potential including the vector projections illustrating the vortex configuration.

Table 3: Parameters describing the asymmetric cases analyzed in this study. The parameters are γ, Eq. 5, softness asymmetry Eq. 6. The distance at the minimum of the PMF is ˜dm.

Asymmetric Pair γ SAB d˜m(˚A) r˜aB,m(˚A) r˜aA,m(˚A)

˜

dm ˜

ra

A,m+˜raB,m

Au140(SC9)62 - Au1289(SC19)258 0.535 0.635 37.5 11.90 25.20 1.011

Au140(SC9)62 - Au1289(SC12)258 0.605 0.866 36.6 11.90 24.75 0.999

Au140(SC9)62 - Au201(SC19)80 0.701 0.576 27.3 11.90 13.80 1.062

Au140(SC19)80 - Au1289(SC9)258 0.803 1.560 38.3 12.25 24.50 1.042

Au1289(SC12)258 - Au1289(SC19)258 0.884 0.733 50.3 24.75 25.20 0.999

Au201(SC19)80 - Au1289(SC9)258 0.928 1.869 38.8 13.80 24.50 1.013

[image:12.612.76.543.533.660.2]
(13)

Asymmetric Nanocrystals

Here we consider two NCs with different radii rA, rB, such that rB < rA. These radii are

given by the OPM formula Eq. 2, which is the equivalent HS radius for NCs, as discussed

further above. Distances corresponding to the minima of calculated asymmetric PMFs will be

denoted with the superscript ”a” while keeping all other notation the same as the symmetric

case i.e. r˜i,m →r˜ai,m. The parameter

γ = rB rA

<1 (5)

is very important, as the equilibrium BNSLs22–24,26are mostly determined by this parameter.

Another important parameter is the softness asymmetry,24 defined as

SAB =

rH B −RB

rH A −RA

, (6)

note that SAB can be larger or smaller than one. A summary of all the cases considered is

given in Table. 3 showing that our study explores γ values in the range [0.535,1.0] and SAB

[image:13.612.85.526.473.643.2]

values between [0.58,1.9].

Figure 6: Representative examples of calculated PMFs show good agreement with a naive prediction based on the sum of symmetric cases.

(14)
[image:14.612.65.515.109.373.2]

Figure 7: Au201(SC9)80NC (purple) and Au140(SC12)62NC (gray) atFmin. There is negligible

interdigitation, and clear vortices and planar interaction surfaces are apparent.

Figure 8: Surface projections of two distinct asymmetric cases, where SAB < 1 (left) and

SAB > 1 (right). The surface distributions shift in location relative to the corresponding

surfaces for the symmetric case of the same particle. This shift always occurs towards large ¯

[image:14.612.98.516.463.613.2]
(15)

the remaining cases shown in the SI. Due to the lack of chain interdigitation, the mid-plane

between two NCs in a symmetric pair defines the interaction plane where space left(right)

of the plane contains all the chains of the left(right) NC. For an asymmetric NC pair, there

is still negligible interdigitation, but the interaction plane requires a precise definition. We

calculate this plane by considering an arbitrary C-united atom in the A-NC and computing

the closest C-united atom in the B-NC, and vice-versa. Iterating over all C-united atoms

in this way, a sub-group of C-united atoms can be selected for each NC (CA, CB). For each

unique atom inCA, the vector joining it to the center of the A NC is computed and projected

along the line connecting the A and B centers. We then used all these projections to define

a histogram that is basically Gaussian. The average of this Gaussian µ effectively defines

the location of the interaction plane, and its standard deviation σ gives an idea of plane

[image:15.612.95.525.371.540.2]

roughness or curvature.

Figure 9: Histograms defining the interaction plane for the Au140(SC9)62NC interacting with

Au140(SC9)62(cyan), Au1289(SC12)258(red), Au201(SC19)80 (blue), Au1289(SC19)258(green).

The location and roughness of the interaction plane are realized in the surface histograms as the mean and standard deviation, respectively. SAB decreases from the top left to bottom

left cases, circling clockwise.

The location of the interaction plane is not a function of the NC radius rA, rB, but

rather, is a function of the softness asymmetry SAB. Vortex textures can be realized as a

(16)

vectors experience an increase in curvature. The location of this structural curvature change

along the ligand chain determines the location of the interaction plane relative to the NC

grafting surface. In general, chains on the ”softer” NC show outward movement of curvature

location along the chain, while chains in the ”harder” NC experience an opposite effect. This

is illustrated in Fig. 8, where relative to the symmetric case, the plane of interaction moves

away from the ”softer” NC, irrespective of which NC has the largest radius. The softness

asymmetry also correlates well with the width, as shown in Fig. 9, where we compare four

cases of an asymmetric pair interaction containing an Au140(SC9)62 NC. These cases show

increasing σ values corresponding with SAB = 1.0,0.866,0.635,0.576, again, irrespective of

which NC has the largest radii. These considerations show that the “atomic orbitals” for

asymmetric NCs are vortices whose detailed structure is a function of the softness asymmetry

SAB.

Discussion

0

1

2

3

λ

(L/R)

1

.

0

1

.

5

2

.

0

2

.

5

τ

(R

min

/D)

1+λ OPM OCM

Au1289(SC9)258

Au1289(SC12)258

Au1289(SC19)258

Au201(SC9)80

Au201(SC12)80

Au201(SC19)80

Au140(SC9)62

Au140(SC12)62

[image:16.612.132.480.428.627.2]

Au140(SC19)62

(17)

General Considerations

The location of the minima of the PMF, 2˜ri,mfor symmetric NC, is expected to be described

by the OCM formula, but instead, a clear breakdown is observed for large curvature, that

is λ > 1, as evident from Fig. 10. Interestingly, as Table 1 shows, the distance is very

sensitive to the NC core radius, but quite insensitive to the chain length L. This is because

for sufficiently large curvature, the chains within the inner OCM cone (see previous section),

extend beyond the actual OCM cone, and therefore the melt condition is not satisfied. A

formula for 2˜ri,m can be elucidated based on our observations, where the separation is given

by the adjusted OCM formula

τA−OCM =

Ri+ ∆i,m

Ri

= 1 + ∆i,m Ri

(7)

where ∆i,m = ri,m − Ri is directly determined by our simulations. The function ∆i,m is

related to the location of gauche chain configurations (broken dihedrals), with values of 5, 7

[image:17.612.77.541.526.653.2]

and 9 ˚A for Au140(SCm)62, Au201(SCm)80 and Au1289(SCm)258, respectively.

Table 4: Predicted values of ˜dm are not simply the sum of symmetric midpoint values, but

correspond to differences in softness. In all cases, the softer NC of the pair increases its asymmetric radius while the harder NC decreases its asymmetric radius, where asymmetric radii are calculated as averages of surface distributions. Rows shaded gray correspond to cases where NC B is softer than NC A.

Asymmetric Pair r˜B,m(˚A)˜rB,ma (˚A)

˜

ra

B,m−r˜B,m ˜

rB,m r˜A,m(˚A) ˜r

a A,m(˚A)

˜

ra

A,m−˜rA,m ˜

rA,m

Au140(SC9)62 - Au1289(SC19)258 11.90 9.32 -0.22 25.20 28.2 0.12

Au140(SC9)62 - Au1289(SC12)258 11.90 10.9 -0.08 24.75 25.7 0.04

Au140(SC9)62 - Au201(SC19)80 11.90 10.1 -0.15 13.80 17.2 0.25

Au140(SC19)80 - Au1289(SC9)258 12.25 14.4 0.18 24.50 23.9 -0.02

Au1289(SC12)258 - Au1289(SC19)258 24.75 23.4 -0.05 25.20 27.0 0.07

Au201(SC19)80 - Au1289(SC9)258 13.80 15.8 0.14 24.50 23.0 -0.06

(18)
[image:18.612.71.548.181.305.2]

Table 5: Free Energy values at the PMF minimum. ∆F = Fmina −

1

2(Fmin,A +Fmin,B) is

always positive, indicating that the system prefers like particle interactions over asymmetric ones in the pair case. Results are given in units ofkBT where T =Tm.

Asymmetric Pair Fa

min Fmin,A Fmin,B ∆F F∆aF

min

Au140(SC9)62 - Au1289(SC19)258 -141 -77 -289 42.0 0.30

Au140(SC9)62 - Au1289(SC12)258 -96 -77 -148 16.5 0.17

Au140(SC9)62 - Au201(SC19)80 -115 -77 -207 27.0 0.23

Au140(SC19)80 - Au1289(SC9)258 -120 -181 -88 14.5 0.12

Au1289(SC12)258 - Au1289(SC19)258 -213 -148 -289 5.5 0.03

Au201(SC19)80 - Au1289(SC9)258 -136 -207 -88 11.5 0.08

Au201(SC9)80 - Au140(SC12)62 -87 -75 -115 8.0 0.09

[image:18.612.190.424.420.598.2]

.

(19)

For asymmetric NCs the minimum of the PMFdmis well predicted by the simple formula:

dm = ˜rA,m+ ˜rB,m, (8)

shown clearly in Table 3. Note, however, that the location of the interacting plane is not a

function of the radius, but rather, is a function of the softness asymmetry, as is evident in

Table 4. We propose a mechanism that drives the plane location change, shown in Figure 11,

based on solving for a plane of interaction with equal contributed area from each NC. This

is an indication of the system’s preference for the symmetric case, supported by the free

energy information in Table 5. We would expect this mechanism to have some degree of

universality, so that any sufficiently short organic ligand would act in a similar manner. The

OTM model supports this claim, as it makes no assumption of ligand type and shows good

agreement with both hydrocarbon and polymer based experimental systems.26

Additional free energy contributions

There are a number of effects that have been neglected in our calculations:

• Core-Core VdW forces.

• Dipole moments.

• Finite cut-off in simulations.

The core-core VdW energy is given by28,39

Uc−c(ρ) =−

AH

6

2RARB

ρ2(R

A+RB)2

+ 2RARB

ρ2(R

A−RB)2

+ log

ρ2(R

A+RB)2

ρ2(R

A−RB)2

(9)

where the Hamaker constant for gold is AH = (3− 5) ·10−19J,39 see Table 11.2. The

calculations shown in Table 6 are computed with the value AH = 5·10−19J. Note that the

(20)

but typically less, rendering our conclusions valid in these cases also. There is usually an

attenuation of Hamaker theory with solvent,39 but solvent effects will not be considered in

this study.

The dipole-dipole moment is given by elementary electrostatics

Ud−d(ρ) =−

~pA·~pB−3(~pA·~n)(~pB·~n)

4π0ρ3

, (10)

where~nis the unit vector joining the centers of A and B. The dipole moment of a CdSe and

PbSe NC is estimated to be of the order of p ≈100D.40–42 Hopefully, ab-initio calculations

will be able to provide a more rigorous estimate for this number. For the calculations shown

in Table 6 we assume a constant dipole density of 100D in a sphere of diameter 5.5 nm,

consistent with Ref.42

All our simulations were conducted with a finite cut-off rcut. A cut-off free result can be

obtained by adding the correction

Ucut(ρ) =

Z

d3rd3r0

ϑ(|r−r0| −rcut)VLJ(|r−r0|)ρH(r)ρH(r0) (11)

where ϑ(x) is the Heaviside function, VLJ is the Lennard-Jones potential between two

hy-drocarbon groups and ρH(r) is the hydrocarbon equilibrium number density.

The magnitude of the different contributions from neglected effects is summarized in

Table 6, which clearly shows none of the effects modify any of the previous conclusions.

An order of magnitude for the magnitude of the Van Der Waals attraction is obtained by

computing the attraction between two planes of area Ac separated a distance D.

U =−Ah

12π Ac

D2 ≈

Ah

12π

(2˜rmtan(χOCM))2

D2 ≈

Ah

12π

(2˜rm)2

D2 (12)

Here, we assumed that the area of contact is 4˜r2tan(χ

(21)

over the range of systems studied, which are of the same order of magnitude as the values

listed in Figure 3 for the internal energy, up to a factor of ≈ 4 or 5 that we attribute to

neglected roughness and boundary contributions.

Another possible effect is the reduction in configuration space of hydrogen motion as

the chains form vortices. This effect is ignored in our calculations, as our simulation used

a united model, but might enhance the entropic repulsion. Without a model describing

explicit hydrogens, this effect cannot be properly estimated. However, we argue that all

chains within a given vortex are separated sufficiently, so that motion of hydrogen atoms is

not significantly hampered and the effect should be small. More detailed atomistic models

[image:21.612.77.540.366.517.2]

will hopefully address this issue.

Table 6: Corrections to free energy by the different contributions: c-c Eq. 9, d-d Eq. 10, cut Eq. 11 in units of kBTm and in % relative to Fmin.

NC Fmin c-c d-d rcut c-c % d-d % rcut %

Au140(SC9)62 -77 -2.3 ±0.03 -2 3.0 0.04 2.6

Au140(SC12)62 -115 -2.4 ±0.03 -4 2.1 0.03 3.5

Au140(SC19)62 -181 -1.7 ±0.03 -8 0.9 0.02 4.4

Au201(SC9)80 -75 -1.2 ±0.03 -3 1.6 0.04 4.0

Au201(SC12)80 -118 -1.5 ±0.04 -5 1.3 0.03 4.2

Au201(SC19)80 -207 -1.3 ±0.03 -12 0.6 0.02 5.8

Au1289(SC9)258 -88 -4.4 ±0.38 -5 5.0 0.43 5.7

Au1289(SC12)258 -148 -3.9 ±0.33 -8 2.6 0.22 5.4

Au1289(SC19)258 -289 -3.2 ±0.33 -21 1.1 0.11 7.3

Conclusions

The main result of this paper is the identification of vortex configurations, see Figs. 5 and 7,

as characteristic ligand structures that determine the bonding between two NCs. In this way,

these vortices play the same role as “atomic orbitals” in determining the covalent bonding

(22)

different sizes through a single parameter, the softness asymmetry Eq. 6, in such a way that

for the softer NC in an asymmetric pair, ˜ra

i,m > r˜i,m. The equilibrium separation distance

is consistent with the OCM formula Eq. 3 for small curvature λ < 1 only, but for large

curvature λ >1 the data is consistent with a formula based on the positions of the broken

dihedrals, independent of ligand length Eq. 7.

The binding free energies are very strong, of the order of several hundreds of kBT with

a significant energy-entropy cancellation, and other effects, namely Van der Waals forces

between cores or dipole-dipole interactions become negligible at this scale, see Table 6. Thus,

the binding of two NCs is determined by the structure of the ligands, that is the vortices,

in complete agreement with the predictions of the OTM.22,26 In extending these results to

BNSLs, it should be noted that the OTM shows that equilibrium distances are dependent

on coordination in the local environment. Therefore, the binding energy of two NCs within

a superlattice is generally lower than that for two isolated ones, but still of the order of

several hundred kBT, see Fig. 3. The relative contribution of the other effects, see Table 6

becomes even smaller, as those decrease much faster with increasing separation. Although

the simulations presented have assumed cores consisting of gold atoms, Table 6 shows that

the validity of our results extends to any type of core, regardless of chemical composition or

physical property, whether insulator, semiconductor or conductor.

Solvent effects have been ignored in our calculations. This is because they are mostly

irrelevant for equilibrium, as BNSLs are completely free from solvent. Yet, such effects are

relevant during the dynamics of self-assembly. In systems assembled by solvent evaporation,

for example, the PMF is repulsive in solvent, and only becomes attractive at low solvent

content, a point nicely illustrated in Ref.31 This critical balance between NC repulsion and

attraction driven by solvent concentration needs to be addressed to obtain complete control

of crystal quality during the process of solvent evaporation.

(23)

evap-component lattices are observed. As shown in Table 5, our results indicate that mixing of

two different isolated NCs is not a favorable process. Still, as predicted by the OTM, there

are many other considerations that may favor mixing, which we are currently investigating.

There are a number of relevant experimental parameters that we have not investigated:

chains with different degrees of unsaturation, the low curvature limit λ << 1 and smaller

grafting densities. Based on our detailed calculations, inclusions of those effects are unlikely

to modify any of the conclusions in this paper, although may result in some quantitative

differences that we will report in the near future.

Materials and Methods

Simulations were carried out in parallel on the GPU resources of the XSEDE Comet cluster

at SDSC. MPI was handled using mpi4py43 across 8 GPUs. MD simulations were carried

out with the HOOMD-blue44,45 MD package in the canonical ensemble at a temperature of

Tm = 387K and non-bonded cut-off of 24˚A. During MD simulations, the Au and S atoms

of each nanocrystal were treated as a rigid body by standard HOOMD-blue commands46 to

preserve the sulfur locations at the “pinning sites” first discussed by Luedtke.47

Pair potential simulations adhered to the following protocol. First, two NCs, bonded at

their centers by a harmonic bond, are simulated at an equilibrium bond distance where the

NCs are not able to interact with one another through the non-bonded interaction cutoff.

Then, the equilibrium bond distance is changed in a stepwise manner, decreasing gradually

in distance to cover the entire range of the pair potential we wish to calculate. At each bond

distance, NCs are allowed to equilibrate, and a configuration to be used in future MPI jobs

is outputted. The configurations created during this stepwise process are then used as initial

states for the further calculation of each window, which is done in parallel.

Simulations in parallel undergo a parallel tempering routine that allows for

(24)

landscape. Final data is taken periodically from simulations 10 million timesteps in length.

The PMF is reconstructed by the WHAM method.48 All the simulations, analysis, etc. are

completely reproducible as they are included as part of the HOODLT49 software. Further

details are provided in SI.

Supporting Information Available

Supplementary information contains one file with:

• Description of the force field and other simulation details.

• Example of the WHAM method and reconstruction of PMF.

• Explanation of parallel tempering method.

• Formulas used to compute the internal energy and Entropy.

• All calculated PMF plots.

This material is available free of charge via the Internet at http://pubs.acs.org/.

Acknowledgement

We acknowledge many discussions and clarifications of their work with M. Boles, D. Talapin

and D. Vaknin. G. Galli provided us insights into the magnitude of dipole contributions. This

work used the Extreme Science and Engineering Discovery Environment (XSEDE), which is

supported by National Science Foundation grant number ACI-1548562. Our project within

XSEDE is supported by grant TG-MCB140071. C.W was initially supported by a

DOE-SULI program and later by NSF. He was in charge of running and analyzing simulations,

(25)

de-the polymer work supported by DOE U.S. Department of Energy (DOE), Office of Science,

Basic Energy Sciences, Materials, Materials Science and Engineering Division and performed

at the Ames Laboratory, which is operated for the US DOE by Iowa State University under

contract number DE-AC02-07CH11358. A.T. designed the project, contributed to the

soft-ware HOODLT and his work is supported by NSF, DMR-CMMT 1606336 “CDS&E: Design

Principles for Ordering Nanoparticles into Super-crystals”. AT wants to acknowledge the

Aspen Center for Physics where some of this work was performed, which is supported by

NSF, PHY-1607611

References

1. Whetten, R. L.; Shafigullin, M. N.; Khoury, J. T.; Schaaff, T. G.; Vezmar, I.;

Al-varez, M. M.; Wilkinson, A. Crystal Structures of Molecular Gold Nanocrystal Arrays.

Acc. Chem. Res. 1999,32, 397–406.

2. Nykypanchuk, D.; Maye, M. M.; van der Lelie, D.; Gang, O. DNA-Guided Crystallization

of Colloidal Nanoparticles. Nature 2008, 451, 549–552.

3. Park, S. Y.; Lytton-Jean, A. K. R.; Lee, B.; Weigand, S.; Schatz, G. C.; Mirkin, C. A.

DNA-Programmable Nanoparticle Crystallization. Nature 2008, 451, 553–556.

4. Zhang, H.; Wang, W.; Mallapragada, S.; Travesset, A.; Vaknin, D. Macroscopic and

Tunable Nanoparticle Superlattices. Nanoscale 2017, 9, 164–171.

5. Zhang, H.; Wang, W.; Mallapragada, S.; Travesset, A.; Vaknin, D. Ion-Specific Interfacial

Crystallization of Polymer-Grafted Nanoparticles. J. Phys. Chem. C 2017,121, 15424–

15429.

6. Zhang, H.; Wang, W.; Akinc, M.; Mallapragada, S.; Travesset, A.; Vaknin, D. Assembling

and Ordering Polymer-Grafted Nanoparticles in Three Dimensions.Nanoscale 2017,9,

(26)

7. Shevchenko, E. V.; Talapin, D. V.; O’Brien, S.; Murray, C. B. Polymorphism in AB13

Nanoparticle Superlattices: An Example of Semiconductor-Metal Metamaterials.J. Am.

Chem. Soc. 2005, 127, 8741–8747.

8. Shevchenko, E. V.; Talapin, D. V.; Murray, C. B.; O’Brien, S. Structural

Characteriza-tion of Self-Assembled MultifuncCharacteriza-tional Binary Nanoparticle Superlattices.J. Am. Chem.

Soc. 2006, 128, 3620–3637.

9. Chen, Z.; OBrien, S. Structure Direction of II-VI Semiconductor Quantum Dot Binary

Nanoparticle Superlattices by Tuning Radius Ratio. ACS Nano 2008,2, 1219–1229.

10. Talapin, D. V.; Shevchenko, E. V.; Bodnarchuk, M. I.; Ye, X.; Chen, J.; Murray, C. B.

Quasicrystalline Order in Self-Assembled Binary Nanoparticle Superlattices. Nature

2009,461, 964–967.

11. Bodnarchuk, M. I.; Kovalenko, M. V.; Heiss, W.; Talapin, D. V. Energetic and Entropic

Contributions to Self-Assembly of Binary Nanocrystal Superlattices: Temperature as

the Structure-Directing Factor. J. Am. Chem. Soc. 2010,132, 11967–11977.

12. Evers, W. H.; Nijs, B. D.; Filion, L.; Castillo, S.; Dijkstra, M.; Vanmaekelbergh, D.

Entropy-Driven Formation of Binary Semiconductor-Nanocrystal Superlattices. Nano

Lett. 2010,10, 4235–4241.

13. Ye, X.; Chen, J.; Murray, C. B. Polymorphism in Self-Assembled AB6 Binary

Nanocrys-tal Superlattices. J. Am. Chem. Soc. 2011, 133, 2613–2620.

14. Boneschanscher, M. P.; Evers, W. H.; Qi, W.; Meeldijk, J. D.; Dijkstra, M.;

Van-maekelbergh, D. Electron Tomography Resolves a Novel Crystal Structure in a Binary

Nanocrystal Superlattice. Nano Lett. 2013, 13, 1312–1316.

(27)

16. Yang, Z.; Wei, J.; Bonville, P.; Pileni, M.-P. Beyond Entropy: Magnetic Forces Induce

Formation of Quasicrystalline Structure in Binary Nanocrystal Superlattices. J. Am.

Chem. Soc. 2015, 137, 4487–4493.

17. Ye, X.; Zhu, C.; Ercius, P.; Raja, S. N.; He, B.; Jones, M. R.; Hauwiller, M. R.; Liu, Y.;

Xu, T.; Alivisatos, A. P. Structural Diversity in Binary Superlattices Self-Assembled

from Polymer-Grafted Nanocrystals. Nat. Commun. 2015,6, 10052.

18. Boles, M. A.; Talapin, D. V. Many-Body Effects in Nanocrystal Superlattices: Departure

from Sphere Packing Explains Stability of Binary Phases.J. Am. Chem. Soc.2015,137,

4494–4502.

19. Boles, M. A.; Engel, M.; Talapin, D. V. Self-Assembly of Colloidal Nanocrystals: From

Intricate Structures to Functional Materials. Chem. Rev. 2016,116, 11220–11289.

20. Murray, M. J.; Sanders, J. V. Close-Packed Structures of Spheres of Two Different Sizes

II. The Packing Densities of Likely Arrangements. Philos. Mag. A. 1980, 42, 721–740.

21. Chen, Z.; Moore, J.; Radtke, G.; Sirringhaus, H.; O’Brien, S. Binary Nanoparticle

Super-lattices in the Semiconductor-Semiconductor Systems: CdTe and CdSe. J. Am. Chem.

Soc. 2007, 129, 15702–15709.

22. Travesset, A. Topological Structure Prediction in Binary Nanoparticle Superlattices.Soft

Matter 2017, 13, 147–157.

23. Travesset, A. Binary Nanoparticle Superlattices of Soft-Particle Systems. Proc. Natl.

Acad. Sci. 2015, 112, 9563–9567.

24. Horst, N.; Travesset, A. Prediction of Binary Nanoparticle Superlattices from Soft

Po-tentials. J. Chem. Phys. 2016,144, 014502.

25. Tkachenko, A. V. Generic Phase Diagram of Binary Superlattices.Proc. Natl. Acad. Sci.

(28)

26. Travesset, A. Soft Skyrmions, Spontaneous Valence and Selection Rules in Nanoparticle

Superlattices. ACS Nano 2017, 11, 5375–5382.

27. Landman, U.; Luedtke, W. D. Small is Different: Energetic, Structural, Thermal, and

Mechanical Properties of Passivated Nanocluster Assemblies. Faraday Discuss. 2004,

125, 1–22.

28. Schapotschnikow, P.; Vlugt, T. J. H. Understanding Interactions between Capped

Nanocrystals: Three-Body and Chain Packing Effects. J. Chem. Phys. 2009, 131,

124705.

29. Schapotschnikow, P.; Pool, R.; Vlugt, T. J. H. Coarse-Grained Model for Gold

Nanocrys-tals with an Organic Capping Layer. Mol. Phys.2007, 105, 3177–3184.

30. Schapotschnikow, P.; Pool, R.; Vlugt, T. J. H. Molecular Simulations of Interacting

Nanocrystals. Nano Lett. 2008,8, 2930–2934.

31. Jabes, B. S.; Yadav, H. O. S.; Kumar, S. K.; Chakravarty, C. Fluctuation-Driven

Anisotropy in Effective Pair Interactions between Nanoparticles: Thiolated Gold

Nanoparticles in Ethane. J. Chem. Phys. 2014,141, 154904.

32. Yadav, H. O. S.; Shrivastav, G.; Agarwal, M.; Chakravarty, C. Effective Interactions

between Nanoparticles: Creating Temperature-Independent Solvation Environments for

Self-Assembly. J. Chem. Phys. 2016,144, 244901.

33. Milowska, K. Z.; Stolarczyk, J. K. Role of Ligand-Ligand vs. Core-Core Interactions in

Gold Nanoclusters. Phys. Chem. Chem. Phys. 2016, 18, 12716–12724.

34. Lange, A.; Danecker, F.; Bauer, G.; Gribova, N.; Gross, J. Different Ways of Looking at

the Force between Two Nanocrystals. J. Chem. Phys. 2015, 143, 244115.

(29)

36. Luedtke, W. D.; Landman, U. Structure and Thermodynamics of Self-Assembled

Mono-layers on Gold Nanocrystallites. J. Phys. Chem. B 1998, 102, 6566–6572.

37. Tachibana, M.; Yoshizawa, K.; Ogawa, A.; Fujimoto, H.; Hoffmann, R. Sulfur/Gold

Orbital Interactions which Determine the Structure of Alkanethiolate/Au(111)

Self-Assembled Monolayer Systems. J. Phys. Chem. B 2002, 106, 12727–12736.

38. Hautman, J.; Klein, M. L. Simulation of a Monolayer of Alkyl Thiol Chains. J. Chem.

Phys. 1989, 91, 4994–5001.

39. Israelachvili, J. Intermolecular and Surface Forces; Academic Press: London, 2000.

40. Shim, M.; Guyot-Sionnest, P. Permanent Dipole Moment and Charges in Colloidal

Semi-conductor Quantum Dots. J. Chem. Phys. 1999, 111, 6955–6964.

41. Li, L.-s.; Alivisatos, A. P. Origin and Scaling of the Permanent Dipole Moment in CdSe

Nanorods. Phys. Rev. Lett. 2003,90, 097402.

42. Talapin, D. V.; Shevchenko, E. V.; Murray, C. B.; Titov, A. V.; Kral, P. Dipole-Dipole

Interactions in Nanoparticle Superlattices. Nano Lett. 2007,7, 1213–1219.

43. Dalcin, L.; Paz, R.; Storti, M. MPI for Python. J. Parallel Distrib. Comput. 2005, 65,

1108–1115.

44. Hoomd-Blue. http://codeblue.umich.edu/hoomd-blue/, Accessed on 24-10-2017.

45. Anderson, J. A.; Lorenz, C. D.; Travesset, A. General Purpose Molecular Dynamics

Simulations Fully Implemented on Graphics Processing Units. J. Comput. Phys. 2008,

227, 5342–5359.

46. Nguyen, T. D.; Phillips, C. L.; Anderson, J. A.; Glotzer, S. C. Rigid Body Constraints

Realized in Massively-Parallel Molecular Dynamics on Graphics Processing Units.

(30)

47. Luedtke, W. D.; Landman, U. Structure, Dynamics, and Thermodynamics of Passivated

Gold Nanocrystallites and Their Assemblies. J. Phys. Chem. 1996, 100, 13323–13329.

48. Kumar, S.; Rosenberg, J. M.; Bouzida, D.; Swendsen, R. H.; Kollman, P. A. The

Weighted Histogram Analysis Method for Free-Energy Calculations on Biomolecules.

I. The Method. J. Comput. Chem.1992,13, 1011–1021.

49. Travesset, A. Phase Diagram of Power Law and Lennard-Jones Systems: Crystal Phases.

(31)

Figure

Figure 1: NCs as soft skyrmions: Hedgehog and vortex, with the different parameters Aiand Ri
Figure 2:Nanocrystal cores Au1289(SCj)258 (left), Au201(SCj)80 (bottom right), andAu140(SCj)62 (top right) with Au atoms shown in yellow and S atoms shown in red
Table 1: Core Diameters and Grafting Densities, see the main text for description.
Figure 3: Potential of Mean Force plots for three representative cases. The minimum aspredicted by OPM Eq
+7

References

Related documents

It also supported by students statement from the result of data questionnaire that the musyrifah 100% female students in building 2 of Ma’had Al Jami’ah stated

For more information, see Using Saved Plans for Provisioning Linux Operating System and Oracle VM Server on Bare Metal Servers7.

For each individual whose compensation must be reported on Schedule 3, report compensation from the organization on row (i) and from related organizations, described in

The novel Untouchables is completely public and provides three possible solutions to eliminate social problems and class exploitation.. For example, the conversion

This thesis is mainly relevant for the master’s students and developers who want to build a new acoustic model in their own language by using the CMU-Sphinx toolkit and also wish

The ECtHR’s case -law in this area suggests that States exercise jurisdiction within the meaning of Article 1 ECHR over third-country victims of (corporate-related) human

Objective: This was a prospective analysis of epidural tramadol as a single analgesic agent delivered through intraoperatively placed epidural catheter for postoperative pain

As stated in the Sunway Sukuk agreement, the Issuer shall undertake to purchase the Trust Assets from the Trustee (acting on behalf of the Sukuk holders) at the