Network


Latest external collaboration on country level. Dive into details by clicking on the dots.

Hotspot


Dive into the research topics where Young Jun Kwak is active.

Publication


Featured researches published by Young Jun Kwak.


Metals and Materials International | 2013

Improvement of hydrogen-storage properties of MgH2 by addition of Ni and Ti via reactive mechanical grinding and a rate-controlling step in its dehydriding reaction

Myoung Youp Song; Young Jun Kwak; Seong Ho Lee; Hye Ryoung Park; Byoung-Goan Kim

In a shift from prior work, MgH2, instead of Mg, was used as a starting material in this work. A sample with a composition of 86 wt% MgH2-10 wt% Ni-4 wt% Ti was prepared by reactive mechanical grinding. Activation of the sample was completed after the first hydriding cycle. The effects of reactive mechanical grinding of Mg with Ni and Ti were discussed. The formation of Mg2Ni increased the hydriding and dehydriding rates of the sample. The addition of Ti increased the hydriding rate and greatly increased the dehydriding rate of the sample. The titanium hydride, TiH1.924, was formed during reactive mechanical grinding. This titanium hydride, which is brittle, is thought to help the mixture pulverized by being pulverized during reactive mechanical grinding and further to prevent agglomeration of the magnesium by staying as a hydride among Mg particles. A rate-controlling step for the dehydriding reaction of the hydrided MgH2-10Ni-4Ti was analyzed by using a spherical moving boundary model on an assumption that particles have a spherical shape with a uniform diameter.


Metals and Materials International | 2015

Hydriding and dehydriding rates of Mg, Mg-10TaF5, and Mg-10NbF5 prepared via reactive mechanical grinding

Myoung Youp Song; Young Jun Kwak; Seong Ho Lee; Hye Ryoung Park

In this work, TaF5 and NbF5 were chosen as additives to enhance the hydriding and dehydriding rates of Mg. Mg, Mg-10TaF5, and Mg-10NbF5 samples were prepared by reactive mechanical grinding. The hydriding and dehydriding properties of the samples were then examined. Mg-10TaF5 had the largest amount of hydrogen absorbed for 30 min and the highest initial dehydriding rate after incubation period, followed in order by Mg-10NbF5, and Mg. At 593 K under 12 bar H2 at the first cycle, Mg-10TaF5 absorbed 3.63 wt% H for 5 min and 4.53 wt% H for 30 min. At 593 K under 1.0 bar H2 at the first cycle, Mg-10TaF5 desorbed 0 wt% H for 2.5 min, 0.59 wt% H for 5 min, 3.42 wt% H for 30 min, and 4.24 wt% H for 60 min. The reactive mechanical grinding of Mg with TaF5 or NbF5 is believed to have facilitated the nucleation and to have decreased the diffusion distances of hydrogen atoms. These two effects are believed to have increased the hydriding and dehydriding rates of Mg. The MgF2 and Ta2H formed in Mg-10TaF5, and the MgF2, NbH2, and NbF3 formed in Mg-10NbF5 are considered to have enhanced both of these effects.


Metals and Materials International | 2013

Comparison of hydrogen-storage properties of Mg-14Ni-3Fe2O3-3Ti and Mg-14Ni-2Fe2O3-2Ti-2Fe

Young Jun Kwak; Hye Ryoung Park; Myoung Youp Song

Magnesium with oxides or transition elements prepared by mechanical grinding under H2 (reactive mechanical grinding) showed relatively high hydriding and dehydriding rates when the content of additives was about 20 wt%. Ni, Fe2O3, and Fe were chosen as the oxides or transition elements to be added. Ti was also selected since it was considered to increase the hydriding and dehydriding rates by forming Ti hydride. Samples Mg-14Ni-3Fe2O3-3Ti (Sample A) and Mg-14Ni-2Fe2O3-2Ti-2Fe (Sample B) were prepared by reactive mechanical grinding, and their hydrogen storage properties were compared. The activated Sample A had a little smaller hydriding rate than the activated Sample B, but a higher dehydriding rate than the activated Sample B. Sample A exhibits quite a larger dehydriding rate and quantity of hydrogen desorbed for 60 min than any other Mg-xNi-yFe2O3-zM (M=transition metals) samples. An addition of a relatively larger amount of Ti is considered to lead to quite a high hydriding rate and a high dehydriding rate of Sample A.


Metals and Materials International | 2015

Preparation of Zn(BH4)2 and diborane and hydrogen release properties of Zn(BH4)2+xMgH2 (x=1, 5, 10, and 15)

Young Jun Kwak; Sung Nam Kwon; Myoung Youp Song

Zn(BH4)2 was prepared by milling ZnCl2 and NaBH4 in a planetary ball mill under Ar atmosphere, and Zn(BH4)2+xMgH2 (x=1, 5, 10, and 15) samples were prepared. Diborane (B2H6) and hydrogen release characteristics of the Zn(BH4)2 and Zn(BH4)2+xMgH2 samples were studied. The samples synthesized by milling ZnCl2 and NaBH4 contained Zn(BH4)2 and NaCl, together with small amounts of ZnCl2 and NaBH4. We designated these samples as Zn(BH4)2(+NaCl). The weight loss up to 400 °C of the Zn(BH4)2(+NaCl) sample synthesized by milling 4 h was 11.2 wt%. FT-IR analysis showed that Zn(BH4)2 was formed in the Zn(BH4)2(+NaCl) samples. MgH2 was also milled in a planetary ball mill, and mixed with the Zn(BH4)2(+NaCl) synthesized by milling for 4 h in a mortar and pestle. The weight loss up to 400 °C of Zn(BH4)2(+NaCl)+MgH2 was 8.2 wt%, corresponding to the weight % of diborane and hydrogen released from the Zn(BH4)2(+NaCl)+MgH2 sample, with respect to the sample weight. DTA results of Zn(BH4)2(+NaCl)+xMgH2 showed that the decomposition peak of Zn(BH4)2 was at about 61 °C, and that of MgH2 was at about 370-389 °C.


Bulletin of Materials Science | 2014

Comparison of hydrogen storage properties of pure Mg and milled pure Mg

Myoung Youp Song; Young Jun Kwak; Seong Ho Lee; Hye Ryoung Park

Hydrogen storage properties of pure Mg were studied at 593 K under 12 bar H2. In order to increase the hydriding and dehydriding rates, pure Mg was ground under hydrogen atmosphere (reactive mechanical grinding, RMG) and its hydrogen storage properties were subsequently investigated. Pure Mg absorbed hydrogen very slowly. At the number of cycles (n) of 1, pure Mg absorbed 0.05 wt% H for 5 min, 0.08 wt% H for 10 min and 0.29 wt% H for 60 min at 593 K under 12 bar H2. Activation was completed at the fifth cycle. At n = 6, pure Mg absorbed 1.76 wt% H for 5 min, 2.17 wt% H for 10 min and 3.40 wt% H for 60 min. The activation of pure Mg after RMG was completed at the sixth cycle. At n = 7, pure Mg after RMG absorbed 2.57 wt% H for 5 min, 3.21 wt% H for 10 min and 4.15 wt% H for 60 min.


Korean Journal of Materials Research | 2013

Improvement of Hydrogen Storage Properties of Mg by Addition of NbF 5 via Mechanical Milling under H 2

Young Jun Kwak; Jiyoung Song; Daniel R. Mumm

A 90 wt% Mg-10 wt% NbF5 sample was prepared by mechanical milling under H2 (reactive mechanical grinding). Its hydriding and dehydriding properties were then examined. Activation of the 90 wt% Mg-10 wt% NbF5 sample was not required. At n=1, the sample absorbed 3.11 wt% H for 2.5 min, 3.55 wt% H for 5 min, 3.86 wt% H for 10 min, and 4.23 wt% H for 30 min at 593K under 12 bar H2. At n=1, the sample desorbed 0.17 wt% H for 5 min, 0.74 wt% H for 10 min, 2.03 wt% H for 30 min, and 2.81 wt% H for 60 min at 593K under 1.0 bar H2. The XRD pattern of the 90 wt% Mg-10 wt% NbF5 after reactive mechanical grinding showed Mg, β-MgH2 and small amounts of γ-MgH2, NbH2, MgF2 and NbF3. The XRD pattern of the 90 wt% Mg-10 wt% NbF5 dehydrided at n=3 revealed Mg, β-MgH2, a small amount of MgO and very small amounts of MgH2 and NbH2. The 90 wt% Mg-10 wt% NbF5 had a higher initial hydriding rate and a larger quantity of hydrogen absorbed for 60 min than the 90 wt% Mg-10 wt% MnO and the 90 wt% Mg-10 wt% Fe2O3, which were reported to have quite high hydriding rates and/or dehydriding rates. The 90 wt% Mg-10 wt% NbF5 had a higher initial dehydriding rate (after an incubation period) and a larger quantity of hydrogen desorbed for 60 min than the 90 wt% Mg-10 wt% MnO and the 90 wt% Mg-10 wt% Fe2O3.


Korean Journal of Metals and Materials | 2013

Hydrogen Storage Properties of a Ni and NbF5-added Mg Alloy

Myoung Youp Song; Young Jun Kwak; Seong Ho Lee; Hye Ryoung Park

Samples with a composition of 80 wt% Mg-14 wt% Ni-6 wt% NbF5 (denoted as Mg-14Ni-6NbF5) were prepared by reactive mechanical grinding, and the hydriding and dehydriding properties of the specimens were then examined. The activation of Mg-14Ni-6NbF5 was completed after two hydriding (under 12 bar H2)-dehydriding (in vacuum) cycles. Mg-14Ni-6NbF5 had a hydrogen storage capacity of about 5.5 wt% H. At the number of cycles n = 2, the sample absorbed 4.93 wt% H for 5 min, 5.20 wt% H for 10 min, and 5.48 wt% H for 60 min at 573 K under 12 bar H2, and desorbed 0.58 wt% H for 10 min, 1.52 wt% H for 30 min, and 2.47 wt% H for 60 min at 573 K under 1.0 bar H2. The hydriding rate increased as the temperature increased from 423 Kto 573 K since the effect of acceleration of thermally activated process predominates, and decreased from 573 K to 623 K since the effect of decrease in the driving force for the hydriding reaction predominates. NbF5 formed MgF2 and NbH2 by the reaction with Mg and hydrogen. Mg-14Ni-6NbF5 exhibited a higher hydriding rate than both Mg-10NbF5 and Mg-14Ni-6Fe2O3.


Journal of Nanoscience and Nanotechnology | 2018

Improvement in the Hydrogen-Storage Characteristics of Magnesium Hydride by Grinding with Sodium Alanate and Transition Metals in a Hydrogen Atmosphere

Myoung Youp Song; Young Jun Kwak; Seong Ho Lee

In this work, MgH2 was used as a starting material instead of Mg. The sample was prepared by grinding MgH2 with sodium alanate and transition metals in a hydrogen atmosphere. Its hydriding and dehydriding properties were measured followed by X-ray diffraction (XRD) analyses and observations of its microstructure. Activation was not required for the 86MgH2 + 10Ni + 2NaAlH4 + 2Ti sample. At the first cycle (n = 1), the sample absorbed 4.96, 5.28 and 5.36 wt% H for 10, 15 and 60 min, respectively, at 593 K in 12 bar H2, showing that the sample absorbed quite a large amount of hydrogen for 60 min (nearly 5.5 wt% H). The initial hydriding rate increased as the temperature increased from 423 K to 553 K and decreased from 553 K to 593 K. The sample showed quite high hydriding rates at relatively low temperatures 423 K (at n = 1) and 473 K (at n = 2) in 12 bar H2, compared with those of other metallic element(s) or compound(s)-added Mg or MgH2 alloys, absorbing 2.89 wt% H for 5 min, 2.97 wt% H for 10 min, and 3.31 wt% H for 60 min at 473 K.


Journal of Nanoscience and Nanotechnology | 2018

Development of an Mg-Based Alloy with High Hydriding and Dehydriding Rates and Large Hydrogen Storage Capacity by Adding TaF5

Young Jun Kwak; Seong Ho Lee; Myoung Youp Song

A sample with a composition of 95 wt% Mg + 5 wt% TaF5 (named Mg-5TaF5) was prepared by reactive mechanical grinding. The activation of Mg-5TaF5 was not necessary, and Mg-5TaF5 had an effective hydrogen storage capacity (the quantity of hydrogen absorbed for 60 min) larger than 5 wt%. At the first cycle (n = 1), the sample absorbed 4.50 wt% H for 10 min and 5.06 wt% H for 60 min at 593 K under 12 bar H2. At n = 1, the sample desorbed 1.58 wt% H for 10 min and 4.93 wt% H for 60 min at 593 K under 1.0 bar H2. The Mg-5TaF5 sample dehydrided at n = 3 contained MgF2 and Ta2H. The hydriding-dehydriding cycling of the sample, which forms MgF2 and Ta2H by reaction with hydrogen, is considered to produce defects on the surface of and inside the Mg particles, to create clean surfaces, and to reduce the particle size of Mg, due to the repetition of expansion with hydrogen absorption and contraction with hydrogen release. Mg-5TaF5 had a higher hydriding rate and a higher dehydriding rate after an incubation period and greater quantities of hydrogen absorbed and desorbed for 60 min than Mg-10TaF5, Mg-10MnO, or Mg-10Fe2O3.


Korean Journal of Materials Research | 2014

Hydrogen Absorption at a Low Temperature by MgH 2 after Reactive Mechanical Grinding

Myoung Youp Song; Seong Ho Lee; Young Jun Kwak; Hye Ryoung Park

Pure MgH2 was milled under a hydrogen atmosphere (reactive mechanical grinding, RMG). The hydrogen storage properties of the prepared samples were studied at a relatively low temperature of 423 K and were compared with those of pure Mg. The hydriding rate of the Mg was extremely low (0.0008 wt% H/min at n = 4), and the MgH2 after RMG had higher hydriding rates than that of Mg at 423 K under 12 bar H2. The initial hydriding rate of MgH2 after RMG at 423 K under 12 bar H2 was the highest (0.08 wt% H/min) at n = 2. At n = 2, the MgH2 after RMG absorbed 0.39 wt% H for 5 min, and 1.21 wt% H for 60 min at 423K under 12 bar H2. At 573 K under 12 bar H2, the MgH2 after RMG absorbed 4.86 wt% H for 5 min, and 5.52 wt% H for 60 min at n = 2. At 573 K and 423 K under 1.0 bar H2, the MgH2 after RMG and the Mg did not release hydrogen. The decrease in particle size and creation of defects by reactive mechanical grinding are believed to have led to the increase in the hydriding rate of the MgH2 after RMG at a relatively low temperature of 423 K.

Collaboration


Dive into the Young Jun Kwak's collaboration.

Top Co-Authors

Avatar

Myoung Youp Song

Chonbuk National University

View shared research outputs
Top Co-Authors

Avatar

Hye Ryoung Park

Chonnam National University

View shared research outputs
Top Co-Authors

Avatar

Seong Ho Lee

Chonbuk National University

View shared research outputs
Top Co-Authors

Avatar

Sung Nam Kwon

Chonbuk National University

View shared research outputs
Top Co-Authors

Avatar

Eunho Choi

Chonbuk National University

View shared research outputs
Top Co-Authors

Avatar

Daniel R. Mumm

University of California

View shared research outputs
Top Co-Authors

Avatar
Top Co-Authors

Avatar

Byung-Soo Lee

Chonbuk National University

View shared research outputs
Top Co-Authors

Avatar

Jiyoung Song

University of California

View shared research outputs
Top Co-Authors

Avatar

Hyuk-Ryeol Park

Mokpo National University

View shared research outputs
Researchain Logo
Decentralizing Knowledge