2018
DOI: 10.1021/acs.jpclett.8b01210
|View full text |Cite
|
Sign up to set email alerts
|

The Clathrate–Water Interface Is Oleophilic

Abstract: The slow nucleation of clathrate hydrates is a central challenge for their use in the storage and transportation of natural gas. Molecules that strongly adsorb to the clathrate-water interface decrease the crystal-water surface tension, lowering the barrier for clathrate nucleation. Surfactants are widely used to promote the nucleation and growth of clathrate hydrates. It has been proposed that these amphiphilic molecules bind to the clathrate surface via hydrogen bonding. However, recent studies reveal that P… Show more

Help me understand this report

Search citation statements

Order By: Relevance

Paper Sections

Select...
3
1
1

Citation Types

0
54
0

Year Published

2019
2019
2024
2024

Publication Types

Select...
6
1

Relationship

1
6

Authors

Journals

citations
Cited by 43 publications
(55 citation statements)
references
References 83 publications
0
54
0
Order By: Relevance
“…The mobility of dodecanol at the clathrate–alkane interface, however, is orders of magnitude faster than at the clathrate–water interface, to which dodecanol binds through insertion of the alkyl tail into half-cages exposed at the surface. 68 We find that a dodecanol molecule takes less than 1 ns to move between two consecutive cages at the clathrate–pentane interface, while the same surfactant at the clathrate–water interface does not move between adjacent cages within microsecond simulations. 68 This difference arises from the distinct mode of binding of dodecanol to the clathrate–oil and clathrate–-water interfaces: the alkyl end of the surfactants is adsorbed deeply into the half-cages at the clathrate–water interface, 68 resulting in a high barrier to move from cage to cage.…”
Section: Results and Discussionmentioning
confidence: 76%
See 2 more Smart Citations
“…The mobility of dodecanol at the clathrate–alkane interface, however, is orders of magnitude faster than at the clathrate–water interface, to which dodecanol binds through insertion of the alkyl tail into half-cages exposed at the surface. 68 We find that a dodecanol molecule takes less than 1 ns to move between two consecutive cages at the clathrate–pentane interface, while the same surfactant at the clathrate–water interface does not move between adjacent cages within microsecond simulations. 68 This difference arises from the distinct mode of binding of dodecanol to the clathrate–oil and clathrate–-water interfaces: the alkyl end of the surfactants is adsorbed deeply into the half-cages at the clathrate–water interface, 68 resulting in a high barrier to move from cage to cage.…”
Section: Results and Discussionmentioning
confidence: 76%
“…Surfactants can bind to both clathrate–water 9,12,13,68,83,86,87 and clathrate–oil 21,37,83 interfaces. Binding of surfactants to clathrate–water interfaces promotes the nucleation 68 and could stall the growth 40,88 of clathrate hydrates. Binding of surfactants to clathrate–oil interfaces is used to prevent the agglomeration of clathrate particles and their coalescence with water droplets.…”
Section: Results and Discussionmentioning
confidence: 99%
See 1 more Smart Citation
“…To identify and quantify the surfactant mechanism at the molecular level, we report single surfactant molecule adsorption onto a hydrate seed and the interaction of the surfactant molecule with the capillary liquid bridge using a classical molecular dynamics (MD) simulation, which has been widely and successfully used in previous investigations of hydrates, including AAs. 15,17,21,24,36 In this paper, we used the general Amber force eld 37 to describe three different surfactant molecules: (1) 1-phenylacetic acid, (2) 2-napthylacetic acid and (3) 1pyreneacetic acid. Using this approach, atom types of each surfactant were taken from the AMBER library.…”
Section: Simulation Tool and System Denitionmentioning
confidence: 99%
“…[14][15][16][17][18][19][20][21] In addition, some reports have shown that AAs could also potentially affect hydrate formation and growth processes. [22][23][24][25] To evaluate the effectiveness of AAs, one frequently used method is to measure model (cyclopentane, CyC5) hydrate interparticle cohesive forces in the presence of AAs using an ambient pressure micromechanical force (MMF) apparatus, 26 because the capillary liquid bridge between hydrate particles was proposed to be another essential feature dominating hydrate particle cohesion in oil continuous systems. [27][28][29] Interestingly, one crucial result of MMF studies of surfactants indicated that polynuclear aromatic carboxylic acids can be highly effective in reducing hydrate cohesion forces by up to 87 AE 9% for a four-ring conjugated hydrophobic group, 16,30 thereby subsequently preventing hydrate particles from agglomerating.…”
Section: Introductionmentioning
confidence: 99%