Você está na página 1de 41

Biodiesel production from sunflower oil over MgO/MgAl 


4 nanocatalyst: 
Effect of fuel type on catalyst nanostructure and performance 

Chemical Engineering Faculty, Sahand University of Technology, P.O. Box 51335-1996, Sahand New Town, Tabriz, Iran 
Reactor and Catalysis Research Center (RCRC), Sahand University of Technology, P.O. Box 51335-1996, Sahand New Town, 
Tabriz, Iran 
a r t i c l e i n f o 
Article history: Received 23 September 2016 Received in revised form 20 November 2016 Accepted 19 December 2016 
Keywords: MgO/MgAl 

a b s t r a c t 
In the present research, the effect of different fuels on preparation of MgAl 

spinel,  as  the  catalyst  sup-  port,  via  the  facile  and 
economical combustion synthesis method is investigated. For this purpose, MgAl 

is  synthesized  by  applying  glycerine,  citric  acid,  ethylene  glycol  and  urea.  It  is  then impregnated by MgO as the active 
phase  of  the  biodiesel  production  process.  The  synthesized  nanocatalysts  were  characterized  by  XRD,  FESEM,  EDX-Dot 
mapping,  BET-BJH,  FTIR  and  TGA  analyses.  Catalytic  activity  of  the  prepared  nanocatalysts  is  evaluated  through  the 
transesterification  reaction  for  the  biodiesel  production  at  110 °C, alcohol-to-oil molar ratio of 12, catalyst-in-feed concentration 
of  3wt.%  and  reaction  time  of  3  h.  Due  to  significant  physicochemical  properties  of  the  sample  prepared  by  urea,  it  exhibits 
considerable  capabilities  in  terms  of  biodiesel  production.  Using  the  prepared  sample  by  urea,  an  oil-to-biodiesel conversion of 
95% is achieved, while the sample is properly stable after six reaction replications. 
Ó 2016 Elsevier Ltd. All rights reserved. 
1. Introduction 
Oil-derived  fuels  represent  about 88% of the fuels commonly used to produce energy [1]. However, in recent years, the use of 
fossil  fuels  resources  has  followed  a  decreasing  trend  because  of  the  reduction in oil reserves and also associated problems with 
these fuels such as environmental pollution and global warming due to the production of gases such as SO 




4 Nanocatalyst Combustion synthesis Fuel type Biodiesel Sunflower oil 
the  transesterification  process;  the  compounds  are,  however,  asso- ciated with disadvantages such as sensitivity to free fatty acid 
and  water,  and difficult and expensive procedure especially for the phase separation [10,11]. Homogeneous acid catalysts such as 


SO 

and HCl are also associated with problems such as the necessity to be at elevated temperatures, inordinate time required to 
perform the reaction, corrosion and high separation costs. Thus, and CO 

[2]. Accord- 
in recent years, researchers have used acidic and basic 
heteroge- ingly, researchers have been seeking to find alternative sources 
neous catalysts for biodiesel production [1,12]. The 
solid acid cat- of energy to fossil fuels [3,4]. Repeatedly noticed during the last 
alysts such as ZrO 
2 decade, biodiesel represents one 
of the alternatives. As a biodegradable and eco-friendly fuel, biodiesel is a mixture of fatty acid methyl esters produced from 
renewable sources such as veg- etable oils and animal fats [1,5–7]. 
The  most  common  way  to  produce  biodiesel  is  transesterifica-  tion  and  esterification  reactions of triglycerides and free fatty 
acids  in  the  presence  of  a  short  chain  alcohol.  Further,  acidic  or  basic  cat-  alysts  are  utilized  for  this  process  [8,9].  Such  basic 
homogeneous catalysts as NaOH and KOH are the primarily used catalysts for 
/SO 
2À, 4 
TiO 

/SO 
2À 4 
and WO 

/ZrO 

can  catalyze  both  esterification  and  transesterification  reactions  [13]. 
Despite  the  advantages  of  this class of catalysts (no sensitivity to the pres- ence of water and free fatty acids in reaction mixture), 
the  longer  time  and  higher  temperatures  required  to  perform  the  reaction  introduce  some  problems  for their application. Unlike 
the solid acid catalysts, the solid basic catalysts need milder conditions for the transesterification reaction. 
Many solid basic catalysts such as La 

[14], KBr [7], CaO-ZrO 



[1], CaSn(OH) 
6 [15], Cs-Na 

ZrO 

[9] have been used for biodiesel production. Among these catalysts, MgO and CaO have 
been used 

Corresponding author at: Reactor and Catalysis Research Center, Sahand University of Technology, P.O. Box 51335-1996, 
Sahand New Town, Tabriz, Iran. 
E-mail address: haghighi@sut.ac.ir (M. Haghighi). 
much  more  than  the  others  for  their  reasonable  cost  and  mild  reac-  tion  conditions  required  [16–18].  They  have  been  usually 
loaded on surface supports such as Al 

URL: http://rcrc.sut.ac.ir (M. Haghighi). 
[19,20] and ZSM-5 zeolites [21] where the supports act in the reaction pathway, too. One 
of 
http://dx.doi.org/10.1016/j.enconman.2016.12.048 0196-8904/Ó 2016 Elsevier Ltd. All rights reserved. 


Energy Conversion and Management 134 (2017) 290–300 

Behgam Rahmani Vahid, Mohammad Haghighi 


Contents lists available at ScienceDirect 

Energy Conversion and Management 


journal homepage: www.elsevier.com/locate/enconman 


 
the  most  significant  features  of  such  supports  is  their  appropriate  surface  area  on  which  the  active  phase  is  to  be  distributed. 
There-  fore,  ceramics  can  present  a  perfect  support  for the catalysis of bio- diesel production process, because of their beneficial 
properties such as large surface area and high thermal resistance [22]. Most recently, spinels such as CoFe 



[23], SrAl 



[24], ZnAl 



[25], ZnFe 



[26], Zn 

Cu 
(1Àx) 
Al 



[27] have been used in the biodiesel production process. However, MgAl 



spinel  has  not  been  used  for  biodiesel  production  reaction  yet.  Due  to  this  spinel’s 
suitable  surface  area  and  high  mechanical  and  thermal  resistance  as  well  as  chemical  resistance  [28–30],  it  can  be  seen  as  a 
promising sup- port for the catalysts used for this purpose. 
The  synthesis  method  plays  major  role  in  physicochemical  properties  of synthesized nanocatalysts [31–33]. Several methods 
have been reported for the synthesis of MgAl 



,  such  as  precipita-  tion,  sol-gel,  citrate-nitrate  and  microwave 
irradiation-assisted  combustion  [34].  However,  due  to  the  simplicity  and  the  short  time  required  to synthesize a catalyst of high 
porosity  [34–36],  the  com-  bustion  method  can  be  an  appropriate  method  to  prepare  a  sup-  ported  catalyst  for  the  biodiesel 
production. In our recent study the MgO/MgAl 



was  used  in  biodiesel production reaction for the first time and the fuel to oxidant ratio was optimized during 
the synthesis of MgAl 



[37].  In  addition  to  the amount of fuel, the type of fuel is known to impose significant effects on final 
properties  of  the  synthesized catalyst by the combustion method [38–40]. The present paper investigates the effects of urea, glyc- 
erin, glycine and ethylene glycol (when used as fuels) on the com- bustion synthesis of MgAl 



as the support on which MgO, as active phase, is to be loaded. 
2. Materials 
In this paper, aluminum nitrate (Al (NO 



Á9H 

O; Merck, 99%), magnesium nitrate (Mg (NO 



Á6H 

O;  Merck,  99%)  and  deionized  distillate  water  were  used  as  precursor  for  the  synthesis of 
MgO/ MgAl 



. The urea (CH 



O; Romil, 99.5%), citric acid (C 





; Merck, 99%), glycerol (C 





; Merck, 99%) and ethylene glycol (C 





;  Merck,  99%)  were  used  as  the  fuel  in  the  combustion  syn-  thesis.  Sunflower  oil  (Acid  content <0.3 mg KOH/gr) and 
methanol (CH 

OH; Merck, 99.99%) were used to biodiesel production reaction. 
2.1. Nanocatalysts preparation and procedure 
The synthesis steps followed to prepare MgO/MgAl 



are demonstrated in Fig. 1. The MgAl 



, as the support for preparing MgO/MgAl 



catalyst,  is  synthesized  by  the  combustion  method  using  different  types  of  fuel.  For  this  purpose,  magnesium 
nitrate  and  aluminum  nitrate  precursors  at  the  Mg/Al  molar  ratio  of  0.5  were  dissolved  in  a  certain  amount  of  distilled  water. 
Once  the  nitrate  precursors  were  dissolved,  the  fuel  (urea,  citric  acid,  glyc-  erol,  ethylene  glycol)  was added to the solution at a 
stoichiometric  ratio  of  1.5.  The  solution  was  stirred  on  a  hot plate at ambient tem- perature for 45 min before being heated to 80 
°C,  so  as  to  have  excess  water  evaporated  to  form  a  viscous gel. The prepared gel was placed in a furnace at 350 °C; after a few 
minutes,  it  was  com-  busted  to  form  a  highly  porous  powder  as  the  final  product.  The  synthesized supports by urea, citric acid, 
glycerol and ethylene gly- col were labeled as MgAl 



(U), MgAl 



(CA), MgAl 



(G) and MgAl 



(EG),  respectively.  The  impregnation  method  was  used  to  load  MgO,  as  active  phase,  on  each  supports,  where  the 
magne- sium precursor was dissolved, at MgO-to-MgAl 



ratio  of  10%wt.,  in  deionized  distilled  water.  Then  the  support  was  added 
and  the  mixture  was  heated  at 80 °C for 180 min. The samples were dried on a hot plate and then placed in an oven at 110 °C for 
24 h under 
B. Rahmani Vahid, M. Haghighi / Energy Conversion and Management 134 (2017) 290–300 291 
an air flow. Finally, the samples were calcined in a furnace at 550 °C for 4 h. 
2.2. Nanocatalysts characterization techniques 
XRD,  FESEM,  EDX,  BET-BJH,  FTIR  and  TGA analyses were used to characterize the synthesized nanocatalysts. A D-5000 
diffractome-  ter  (Germany,  Siemens,  Cu  Ka  radiation,  0.154056  nm) was used in the range of h = 10–80° to study the formation 
of  crystals  and  crys-  talline  phase  by  XRD  analysis.  Commonly  used  to  examine  surface  morphology,  FESEM  analysis  was 
undertaken  by  a  MIRA3  FEG-SEM  (Czech  Republic,  TESCAN).  Also,  a  VEGA  II  Detector (Czech Repub- lic, TESCAN) was 
used  to  perform  EDX-Dot  mapping  analysis  to  determine  surface  composition  of  the  synthesized  nanocatalysts.  In  order  to 
examine  surface  area,  pore  diameter  and  pore  volume  of  the  synthesized  samples,  BET-BJH  analysis  was  conducted  on  a 
Chembet-3000  (USA,  QuantaChrome)  analyst.  A  TENSOR  27  (Ger-  many,  Bruker)  analyst  was used to perform FTIR analysis 
in  the  range  of  400–4000  cmÀ1  on  K-Br tablets, so as to detect formed functional groups as well as remaining nitrate precursors 
and  fuel  in  the  samples.  TGA  analysis  was  performed  by  a  PYRIS  Diamond  (USA,  Perkin  Elmer)  in  the  temperature  range  of 
50–700 °C at a temperature rate of 10 °C/min; the analysis were undertaken under an air flow for all samples. 
2.3. Experimental setup for catalytic performance test 
A  stainless  steel  reactor  with  a  100  ml  inner  Teflon  container  was  used  for  the  biodiesel  production  reaction.  A  heat  clamp 
with  a  digital  controller  was  used  to  adjust  the  required  temperature,  with  a  magnetic  stirrer  utilized  to  stir  the  mixture.  The 
reaction  was  carried  out  based  on  the  conditions  of  our  previous  work  [37]:  110  °C,  an  alcohol-to-oil  molar  ratio  of  12, 
catalyst-in-oil  con-  centration  of  3  wt.%,  and  3  h  of  reaction  time.  The  reaction  prod-  ucts  were  analysed  on  a  gas 
chromatography  (GC  Chrom,  Teif  Gostar  Faraz,  Iran) equipped with a FID and a SupraWax-280 (Spain, Technokroma) column, 
so  as  to  evaluate  the  conversion.  1  ml  of  each  sample was injected at split mode (1/100), while the FID and injector temperature 
were set to 260 °C. Also, the standard EN-14103 was used to calculate the amount of FAME in each sample. 
3. Results and discussions 
3.1. Nanocatalysts characterization 
3.1.1. XRD analysis 
The  XRD  analysis  of  the synthesized nanocatalysts is presented in Fig. 2. Compared to other samples, the synthesized sample 
by  urea  had  good  crystallinity.  But  the  samples  prepared  with  citric  acid  and  glycerine  had  low  intensity  peaks  and  almost 
amorphous  structure  due  to  incomplete  combustion  of  the  fuels,  generating  inadequate heat during the combustion. Considering 
the standard JCPDS pattern (Cubic, 00-001-1157) of MgAl 



with  characteristic  peaks  at  2h  =19.2,  31.6,  37.3,  45.3,  60.0  and  66.2,  the 
results proved the formation of MgAl 



crystals  in  all  of the synthesized samples. Also, small peaks at 2h = 37.0, 43.0, 62.4, 74.8 and 
78.7  (corresponding  to  the  standard  pattern  of  MgO  (JCPDS;  Cubic,  01-  077-2364))  for  the  nanocatalysts prepared by urea and 
ethylene glycol indicate smaller size and well dispersion of MgO crystals on the surface of MgAl 



;  no  peaks  were  properly diagnosed for the other samples, probably due to either the absence of MgO 
or amorphous structure of the formed MgO. The characteristic Al 


3  peaks  at  2h  =  37.4,  39.6,  42.8,  45.7,  60.4  and  67.3 
(JCPDS; Cubic, 00-004-0880) were not observed in any of the samples, indicating 
 
the  absence  of  the  oxide  or  lack  of its crystalline phase. Crystal size [41] and relative crystallinity of the synthesized samples are 
pre- sented in Table 1. According to the table, the relative crystallinity and crystal size of MgAl 



(U)  are  higher  than  those  of  the  others,  because  of far better combustion reaction of urea leading to 
the generation of a larger amount of heat. Among other supports, MgAl 



(EG)  had  a  better  crystallinity,  with  the  other  two  samples  having  almost  the  same  crystallinity.  The  lower 
crystallization  degree  and  smaller  crystal  size  of  the  samples  could  be  attributed  to  incomplete  combustion  of  citric  acid  and 
glycerol. 
3.1.2. FESEM analysis 
The  results  of  the  FESEM  analysis  of  the  synthesized  nanocata-  lysts  are  shown  in  Fig.  3.  It  clearly  shows  that  the  sample 
synthe-  sized  by  urea  exhibited  a  honeycomb-shaped  morphology.  With  large  pores  on  its  surface,  this  catalyst  seems  to  be  an 
appropriate  catalyst  for  the  biodiesel  production  reaction  [42,43]. Morpholo- gies of the samples synthesized with citric acid and 
glycerin  reflected  amorphous  structure  of  the  catalysts,  as  it  was  also  indi-  cated  by  the  results  of  the  XRD  analysis;  these 
samples  have  quite  different  appearances  from  that  of  the  former  sample  (honeycomb-  shaped).  It  can be due to the incomplete 
combustion  reaction  and  subsequently  low  heat  generated  by  those  fuels,  as  mentioned  in  the  section  on  XRD  analysis. 
Appearance  and  morphology  of  the  synthesized  catalyst  with  ethylene  glycol  were  somewhat  similar  to  those  of  the  sample 
synthesized by urea, except that agglomer- 
292 B. Rahmani Vahid, M. Haghighi / Energy Conversion and Management 134 (2017) 290–300 
(a) Precursors preparation 
Fuel 
Precursors 
Urea Citric Acid 
Mg(NO 



.6H 


Al(NO 



.9H 


Glycerol Ethylene Glycol 
Aq. solution of Mg and Al nitrates in appropriate ratios (1/2 molar ratio) 
Mixing for 45 min with fuel ratio of 1.5 
(b) Combustion synthesis of support 
Heating over a hot plate at 80°C to evaporate excess water and form viscose gel 
Heating in a furnace maintained at 350oC 
Gel boiling with frothing and foaming to form combustion powder 
(c) Impregnation synthesis of nanocatalyst 
Mg(NO 
3 MgAl 



(G), MgAl 
2 Support: O 

(CA), MgAl 
MgAl 2 


O 4 

(EG), MgAl 



(U) 
Mg Precursor 


.6H 

O De-ionized water 
Impregnation at 80°C for 180 min 
(d) Post treatment of nanocatalyst 
Drying at 110°C for 24 h in air ambient 
Calcination at 550°C for 4 h in air ambient 
MgO/MgAl 



(G), MgO/MgAl 
Nanocatalyst: 2 O 

(CA), MgO/MgAl 
MgO/MgAl 2 


O 4 

(EG), MgO/MgAl 



(U) 
Fig. 1. Combustion synthesis of MgO/MgAl 



nanocatalyst using various fuels. 
ated  particles  could  be  seen  clearly,  even  though  they  had  not  inhibited  the  formation  of  large pores through which triglyceride 
molecules  could  pass.  It  seems  that,  considering  the  large  size  of  triglyceride  molecules  and  associated mass transfer problems, 
the  synthesized  catalysts  with  urea  and  ethylene glycol had suit- able morphologies for being used in transesterification reaction. 
Fig.  4  shows  the  distribution  of  surface  particle  sizes  [44]  for  the  prepared  catalyst  with  urea  as  fuel.  Minimum  and  maximum 
par-  ticle  sizes  were  5.7  and  20.1  nm,  respectively,  with  an  average  size  of  11  nm that this results were in good agreement with 
the results of XRD analysis. 
3.1.3. EDX analysis 
The  transesterification  reaction  of  triglyceride  for  biodiesel  pro-  duction  is  a  surface  reaction  at  catalytically  active  sites  on 
MgO. So in this case, the dispersion of active sites on the support surface plays an important role which can be evaluated by EDX 
analysis [45,46]. The EDX analysis of MgO/MgAl 



nanocatalysts  synthe-  sized  with different types of fuel are presented in Fig. 5. Dot 
map-  ping  images  delineated  that  Mg  and  Al  elements  were almost uniformly distributed for all samples, with no agglomeration 
of  elements  being  seen,  which was an important parameter when evaluating the catalysts [47]. The issue was also understandable 
in  terms  of  the  measured  percent  of  each  element  by  the  EDX  anal-  ysis, as is shown in Fig. 6. In this figure, the percentages of 
Mg, Al and O elements as well as their theoretical percentages (the per- 
 
not  formed or a small amount of the synthesized catalysts were crystalized by citric acid and glycerol. So, the samples had almost 
amorphous  structures  with  finer  particles  from  the  crystalline  phases,  presumably  accounting  for  their  large  surface  areas  and 
pore  volumes.  But  the  synthesized  samples  with  citric  acid  and  glycerol  had  a  pore  diameter  slightly  less  than  the  synthesized 
samples  by  either  urea  or  ethylene  glycol,  possibly  due  to  incom-  plete  combustion  and  low  exhaust  gases  derived  from  the 
consid- ered samples. Comparing mean pore diameter of the MgO/MgAl 


4  (EG)  sample  and  the  large  size  of  the  triglycerides 
molecules,  it  can  be  concluded  that  this  catalyst  can  be  suitable  for  biodiesel  pro- duction [42,43]. However, the surface area of 
this  catalyst  was  much  smaller  than  those  of  the  others,  imposing  a  negative  effect on the reaction. Among the four synthesized 
catalysts, the MgO/ MgAl 



(U)  had  an  appropriate  surface  area and mean pore diam- eter, indicating better activity of the catalyst in comparison to 
other  catalysts  in  the  reaction.  The  adsorption-desorption  hystere-  sis  and  pore  size  distribution  plots  of  the  catalysts  are 
illustrated  in Fig. 7. According to the figure, the largest pore size distributions of the synthesized samples with urea, glycerin, and 
citric  acid  were  almost  identical and within the range of 2–8nm that was an appropriate range for heterogeneous catalysts used in 
the  produc-  tion  of  biodiesel.  The  synthesized sample by ethylene glycol had a greater pore size distribution in the range of 6–14 
nm;  this  differ-  ence  could  be  due  to  higher  gas  generation  in  more  complete  com-  bustion  than  the  other  fuels.  The 
adsorption/desorption  diagrams of the samples showed IV type based on IUPAC classification, cor- responding to a material with 
pore diameters of 1.5–100 nm. The hysteresis plots of MgAl 



(CA) and MgAl 



(G)  samples  are  of  type  H3  (slit-shaped  pores),  making  them  not  suitable  for 
the bio- diesel production reaction. The MgAl 

(U)  exhibit  hysteresis  of  types  H1  (cylindrical  pore  channels) 
and H2 (disordered pores), respectively, that are more suitable structures for biodiesel production, apparently. 
3.1.5. FTIR analysis 
The  FTIR  analysis  of  the  synthesized  catalysts  is  shown  in  Fig. 8. Significant peaks in the range of 1640 and 3500 cmÀ1 are 
attribu-  ted  to  O–H  stretching  and  bending  of water adsorbed onto the sur- face of the synthesized catalyst, respectively [48–51]. 
The peaks around 2930 and 2860cmÀ1 are appointed to C–H stretching vibrations bond of fuels used in the combustion synthesis 
[29,52].  The  peak  at  1100  cmÀ1  is  also  related  to  the  C–N  group  of  the  fuels  [53].  The  corresponding  peaks  to  C–H and C–N 
indi-  cated  that  some  amounts  of  the  fuel  remained  in  all  catalysts,  so  that  the  exact  amount  of  the  fuel  cannot  be  specified. 
Belonging to vibrate bonds of NO 
3 MgO MgAl 



(a) MgO/MgAl 



(G) 
(b) MgO/MgAl 



(CA) 
(c) MgO/MgAl 



(EG) 
Reference patterns 
(d) MgO/MgAl 



(U) 
MgO (Cubic, 01-077-2364) 
MgAl 



(Cubic, 00-001-1157) 
Al 



(Cubic, 00-004-0880) 
10 15 20 25 30 35 40 45 50 55 60 65 70 75 80 85 90 2 (degree) 


(EG) and MgAl 



Fig. 2. XRD patterns of combustion synthesized MgO/MgAl 



nanocatalysts using various fuels: (a) MgO/MgAl 



(G), (b) MgO/MgAl 



(CA), (c) MgO/MgAl 



(EG), and (d) MgO/MgAl 



(U). 
centage of the initial gel of MgAl 



support  and  the  percentage  of  elements  loaded  on  the  support  during  the  impregnation 
process)  are  provided  for  each  catalyst.  Although  the percentages of the ele- ments were in a relatively good agreement with one 
another, the small variance in between them could be due to local analysis of the catalyst. 
3.1.4. BET-BJH analysis 
To obtain the diameter, pore volume and surface area of the synthesized samples, BET-BJH analysis were carried out; the 
À groups [52], the peaks at 
1520, 1400 and results are shown in Table 1 and Fig. 7. Among the synthesized 
850 cmÀ1 indicated that some amount of nitrate in 
precursors is samples, MgAl 



(CA) and MgAl 



(G) represent much larger 
probably remained in the catalyst support. However, 
TG analysis surface area and pore volume than the MgAl 



(U) and MgAl 



was required to evaluate the remained amounts of 
nitrate and fuel (EG) samples. The large difference might be because of their mor- 
in the samples (see the section on TG analysis). The 
peaks at 470, phology and crystallinity, as mentioned in the XRD and FE-SEM 
520 and 710 cmÀ1 came from AlO 

groups as well as 
lattice vibra- sections. It can be explained in a way that the crystals either were 
tion of Mg–O stretching, representing successful formation of the 
Table 1 Structural properties of synthesized MgO/MgAl 

nanocatalyst using various fuels. 
Nanocatalyst MgO 
(wt.%) 


MgAl 



Fuel 
Fuel type Surface (wt.%) 
ratio 
area (m2/g) 
Pore volume 
Mean pore 
Relative crystallinitya Crystallite sizeb (nm) (cm3/g) 
size (nm) 
MgAl 


MgO/MgAl 



MgAl 





(G) 10 90 1.5 Glycerol 111.8 0.1534 6 13.1 2.4 MgO/MgAl 



(CA) 10 90 1.5 Citric acid 184.1 0.3579 5.1 21.4 3.0 MgO/MgAl 



(EG) 10 90 1.5 Ethylene glycol 34.1 0.1102 8.8 37.2 7.9 MgO/MgAl 



(U) 10 90 1.5 Urea 60.6 0.0787 6.3 100 21.3 
a Relative crystallinity: XRD relative peak intensity at 2h = 66.2°. b Crystallite size estimated by Scherre’s equation at 2h = 
66.2°. c Crystallite phase: Cubic (JCPDS: 00-001-1157, 2h = 19.2, 31.6, 37.3, 45.3, 60.0 and 66.2). 
B. Rahmani Vahid, M. Haghighi / Energy Conversion and Management 134 (2017) 290–300 293 
 
spinel MgAl 

[29,53,54]. These results were consistent with other analyses, especially the XRD analysis. 
3.1.6. TG analysis 
TG  analysis  of  the  synthesized  samples  using  different  types  of  fuel  is  shown  in  Fig.  9  for  detection  of  remaining  fuel  and 
precur-  sors  in  the  samples.  The  reduction  in  the  range  of  100–250  °C  in  all  of  the  samples  is  related  to  adsorbed water on the 
catalyst  sur-  face.  The  remained  precursor  nitrates  were  removed  in  the  range of 250–400 °C where various reductions could be 
observed for differ- ent samples. MgAl 



(G) and MgAl 



(CA) catalysts were observed with about 4% and 11% of reductions, respectively. 
Partial 
(G) 
(b) MgO/MgAl 



(CA) 
(c) MgO/MgAl 



(EG) 
(d) MgO/MgAl 



(U) 
Fig. 3. FESEM images of combustion synthesized MgO/MgAl 

(
EG), and (d) MgO/MgAl 



nanocatalysts using various fuels: (a) MgO/MgAl 


4 O 

(U). 
294 B. Rahmani Vahid, M. Haghighi / Energy Conversion and Management 134 (2017) 290–300 


(a) MgO/MgAl 

decomposition  of  the  precursor  nitrates  could  be  due  to  the  low  heat  caused  by  incomplete  combustion. About 2% of reduction 
could be seen in the MgAl 



(EG),  indicating  larger  amount  of  heat  generated  during the synthesis of these samples rather than 
the former two samples. The MgAl 



(U) apperceived no significant decrease, representing the appropriate temperature during the 
combustion  leading  to  complete  decomposition  of  the  nitrates.  At  temperatures from 400 to 700 °C, for the samples synthesized 
by  glycerin,  citric  acid  and ethylene glycol as fuel, about 4, 3 and 3% of mass reductions were observed, respectively. This could 
be  ded-  icated  to  the  remaining  fuel  or  created  carbons  from  incomplete  combustion  of  the  corresponding  fuel.  This  mass 
reduction for 


(G), (b) MgO/MgAl 



(CA), (c) MgO/MgAl 



 
these  samples  proved  that  the  combustion  has  not  occurred  thor-  oughly,  so  that  the  required  heat  for  the  decomposition of the 
nitrates was not provided. There was no mass reduction in the MgAl 



(U)  for  the  same  temperature  range  and  as  mentioned  before; it could be due to the complete combustion of the corre- 
sponding fuel during the synthesis of the sample. 
3.2. Catalytic performance study toward biodiesel production 
One  of  the  most  important  criteria  for  a  catalyst  is  its  perfor-  mance  in  a  reaction.  Therefore,  as  a  final  analysis,  the 
synthesized  catalysts  were  evaluated  in  the  transesterification reaction of sun- flower oil. All reactions were performed under the 
same  conditions  as  described  in  experimental  setup  section.  Process  conditions  for  the  reaction  have  great  impact  on  the  final 
conversion  while  they  also  play  a  key  role  in  determination  of  the  cost  of  reaction.  It indi- cated that the best catalyst is the one 
which  converts  more  even  in  milder  reaction  conditions.  Accordingly,  in  this  research,  the  pro-  cess  conditions  were  chosen 
somehow  to  make  it  possible  to  com-  pare  the  results  with  other  researchers’  studies.  The  reaction  products  were  tested  once 
separated  from  the  produced  glycerin by GC analysis according to the EN-14103 standard method, with the results shown in Fig. 
10.  As  was  expected  based  on  the  results  of  the  other  analyses,  the  highest  conversion  (about  95.7%)  was  related  to  the 
MgO/MgAl 



(U) catalyst. The MgO/MgAl 



(EG)  achieved  a  conversion  of  84.5%,  i.e.  around  10%  lower 
than  that  of  the  former  sample. As explained in previous sections, the supe- rior conversion of this sample was predictable due to 
its large pore 
MgO/MgAl 



(U) 
100 
Min: 5.7 nm Max: 20.1 nm 80 
Average = 11.0 nm 
20 nm 
60 
52.3 

50nm 40 
16.4 
6.3 

5-10 10-15 15-20 20-25 
Particle size (nm) 
Fig. 4. Surface particle size distribution histogram of MgO/MgAl 
2 25.0 
20 
B. Rahmani Vahid, M. Haghighi / Energy Conversion and Management 134 (2017) 290–300 295 
diameter. Moreover, the lower activity of this sample compared to MgO/MgAl 



(U)  could  be  interpreted  by  its  smaller  surface  area  and  lower  crystallinity.  The  corresponding  conversions  to 
MgO/ MgAl 



(CA) and MgO/MgAl 



(G)  catalysts  were  58%  and  51%,  respectively.  The  relatively  low  conversions  could  be 
attributed to poor crystallinity of the catalysts and the formation of amor- phous phase of MgAl 



.  With  regard  to  the  discussion,  it  can be concluded that the large surface area and suitable pore size of 
the MgAl 



(when  used  as  catalyst  support)  cannot  solitarily  determine  the  best  sample.  If  relative  crystallinity  is  high,  pore 
diameter and surface area can be considered as two key factors for biodiesel production reaction. The conversion of MgO/MgAl 


4 (U) (the best catalyst in the samples examined in the 
present work) was comparable to the reported results by researchers who had used Mg-Al mixed oxide catalyst which showed 
high conversions in milder reaction conditions. C ˇ 
apek  et  al.  as  well  as  Hájek  et  al.  used  Mg-Al  mixed  oxide  for  biodiesel  production;  they 
achieved  a  conversion  of  78%  at  117  °C,  an  alcohol-to-oil  molar  ratio  of  24,  a  catalyst-in-oil  concentration  of  4  wt.%,  and  a 
relatively  long  reac-  tion time of 8 h [55,56]. In another work, Brito et al. achieved a conversion of 90% at a reaction temperature 
of  160  °C, an alcohol-to-oil molar ratio of 12, a catalyst concentration of 6 wt. %, and a reaction time of 6 h [57]. In several other 
articles  where  the  Mg-Al  mixed  oxides  were  used,  even  though  under  severer  reaction  conditions,  the  authors  reached 
conversions  lower  than  or  equal  to  those  of  the  present  study  [58–60].  Beside  all  the  fore-  going  advantages,  in  this  study,  the 
selected catalyst was synthe- sized by a far simpler and faster method than all the above 


(U) nanocatalyst. 
 
mentioned catalysts. This can be attributed to the elimination of time wasting and costly calcination and drying process in support 
preparation  procedure.  To  put  it  in  nutshell,  it  can  be  said  that  this  catalyst  has  the  suitable  economic  ability  for  industrial 
production  according  to  its properties, which was concluded from different analysis and also its performances. With regard to the 
above- mentioned hints, the MgO/MgAl 



(U)  catalyst  (the  optimum  cat-  alyst  in  the  present  work) can be said to have good features 
in terms of simple and economical synthesis, making it a suitable heterogeneous catalyst for biodiesel production process. 
AlMg O 
(c) MgO/MgAl 

2.0μm 
AlMg O 
Fig. 5. EDX analysis of combustion synthesized MgO/MgAl 

AlMg O 
2.0μm 
2.0μm 


(EG) 
(d) MgO/MgAl 



(U) 
keV 
keV 
(
EG), and (d) MgO/MgAl 



nanocatalysts using various fuels: (a) MgO/MgAl 



(G), (b) MgO/MgAl 



(CA), (c) MgO/MgAl 


4 O 

(U). 
Mg Al O 
100 
44.0 
36.7 
Parent Composition 
Fig. 6. Parent solution vs. surface chemical analysis of combustion synthesized MgO/MgAl 

19.8 
24.3 25.2 

21.0 ) % t w ( 
24.8 80 
n o i t i s o p m o C 
60 
35.0 
38.6 
31.8 
39.3 
28.3 
40 
20 
48.5 
36.3 
46.5 
MgO/MgAl2O4 
MgO/MgAl2O4 
MgO/MgAl2O4 
MgO/MgAl2O4 (G) 
(CA) 
(EG) 
(U) Nanocatalyst 


nanocatalyst using various fuels. 
296 B. Rahmani Vahid, M. Haghighi / Energy Conversion and Management 134 (2017) 290–300 
O (a) MgO/MgAl 


(G) 
keV 
AlMg O 
2.0μm 
(b) MgO/MgAl 

3.3. Activity performance of nanocatalyst toward biodiesel production 
Reusability  of a catalyst in corresponding processes is a very important parameter in the study of heterogeneous catalysts. The 
reusability  is  measured  by  various  methods,  among  which  one  can  refer  to  the  number  of  rounds  the  catalyst  can  be  used  in  a 
batch reaction. As the optimum catalyst in the present work, MgO/MgAl 



(U)  was  evaluated  at  reaction  conditions  to  deter-  mine  its  reusability.  For  this  purpose,  the  used  catalyst  was 
sepa-  rated  from  the  reaction  products  after  each  round of the reaction; it was then simply washed with methanol and put into an 
oven  at  110  °C  for  24  h  before  being  reused. The sample suc- ceeded to be usable for six successive rounds of biodiesel produc- 
tion  reaction,  with  the  conversion  of  each  round  assessed  by  GC  analysis;  the  results  are  shown  in  Fig.  11.  In  the  figure,  the 
activity  of  each  round  has  been  obtained  by  dividing  the  conversion  of  the round by the conversion of the first round multiplied 
by  100.  According  to  the  figure,  it  is  clear  that  catalytic  activity  reduced  by  about  5%  and  10%  in the second and third rounds, 
respectively.  But  in  further  rounds,  the  catalytic  activity  remained relatively constant, with some negligible reductions. It should 
be  noted  that  one  of the causes of catalyst deactivation in the biodiesel produc- tion process is washing and leaching of the active 
phase from sup- port surface. Therefore, the large reduction in catalyst activity of MgO/MgAl 



(U)  in  the  second  and  third  runs  could  be  related  to  the  leaching  of  some  of  active  phase  in  reaction  mixture 
which  had  weak  bond  with  the  support  surface,  reducing the activity of the catalyst. The washing and leaching of the MgO from 
the surface 


(CA) 
keV 
 
of  the  support  can  be  attributed  to  the  high  viscosity  of  the  oil  and  also  the  high  temperature  and  pressure  of  the  reaction. The 
active  phase  leaching  and  its entrance to the liquid phase of the reaction not only can deactivate the catalyst, but it can also affect 
the  qual-  ity  of  the  produced  biodiesel  fuel.  However,  as  the  remained  active  phase  performed  suitable  interaction  with  the 
support surface, the 
(a) MgO/MgAl 



(G) 100 
0.0 0.2 0.4 0.6 0.8 1.0 Relative Pressure (p/p°) 

0.25 Adsorption Desorption 
0.20 
Desorption 
0.15 Adsorption 
0.10 
0.05 
0.00 
0 2 4 6 8 10 12 14 16 18 
Pore Diameter (nm) 
(b) MgO/MgAl 



(CA) 250 
Adsorption 
0.50 
200 
Desorption 
0.40 Desorption 150 
0.30 
100 
Adsorption 
0.20 
50 
0.10 
0.00 0.0 0.2 0.4 0.6 0.8 1.0 
0 2 4 6 8 10 12 14 16 18 Relative Pressure (p/p°) 
Pore Diameter (nm) 
(c) MgO/MgAl 



(EG) 
) g / 
80 
70 
Adsorption 
0.00 0.0 0.2 0.4 0.6 0.8 1.0 

0.10 
m c ( d 
60 
Desorption 
) g / 
0.08 
e b r 
50 
m c ( ) 
0.06 o s d 
40 

30 
Desorption 
D ( g o l 
0.04 y t i t n a u Q 
20 
10 

Adsorption 
d / V d 
0.02 
0 2 4 6 8 10 12 14 16 18 Relative Pressure (p/p°) 
Pore Diameter (nm) 
(d) MgO/MgAl 



(U) 
) g / 
60 
Adsorption 
0.16 
m c ( 
50 
Desorption 
d e 
40 
) g / 
0.12 
b r 
30 
Desorption o s d A y t 
20 
Adsorption 
m c ( ) D ( g o l 
0.08 
i t n 
10 
d / V 
0.04 a u Q 

0.00 0.0 0.2 0.4 0.6 0.8 1.0 
0 2 4 6 8 10 12 14 16 18 Relative Pressure (p/p°) 
Pore Diameter (nm) 
Fig. 7. Adsorption/desorption isotherms and pore size distribution of combustion synthesized MgO/MgAl 

(
G), (b) MgO/MgAl 



nanocatalysts using various fuels: (a) MgO/MgAl 
2 O 

(CA), (c) MgO/MgAl 



(EG), and (d) MgO/MgAl 



(U). 
B. Rahmani Vahid, M. Haghighi / Energy Conversion and Management 134 (2017) 290–300 297 
80 
60 
40 
20 

active  phase  was  prevented  from  being  easily  leached  during the proceeding rounds (fourth to sixth rounds), so that the catalytic 
activity  remained  almost  constant.  It  is  notable  that after 6 runs, the catalyst’s conversion reached nearly 80% (83% of the initial 
value) which is higher than the reported results of other research- ers that used somehow similar catalysts. 


 
catalyst was synthesized with the simple and applicable method of combustion for being used 
(a) MgO/MgAl 



(G) 
(b) MgO/MgAl 



(CA) 
(c) MgO/MgAl 



(EG) 
(d) MgO/MgAl 



(U) 
4000 3500 3000 2500 2000 1500 1000 500 
Fig. 8. FTIR spectra of combustion synthesized MgO/MgAl 



nanocatalysts using various fuels: (a) MgO/MgAl 



(G), (b) MgO/MgAl 



(CA), (c) MgO/MgAl 



(EG), and (d) MgO/MgAl 



(U). 
100 
98 
(a) MgO/MgAl2O4 (U) 
96 
) % ( s s o l 
94 
92 
(b) MgO/MgAl2O4 (EG) 
t h g i 
90 
(c) MgO/MgAl2O4 (G) e W 
88 
86 
Zone-I 100-250oC 
80 
100 150 200 250 300 350 400 450 500 550 600 650 700 Temperature (°C) 
(d) MgO/MgAl2O4 (CA) 84 
82 
Zone-II 
Zone-III 250-400oC 
400-700oC 
Fig. 9. TG analysis of combustion synthesized MgO/MgAl 



nanocatalysts using various fuels: (a) MgO/MgAl 



(G), (b) MgO/MgAl 



(CA), (c) MgO/MgAl 



(EG), and (d) MgO/MgAl 



(U). 
298 B. Rahmani Vahid, M. Haghighi / Energy Conversion and Management 134 (2017) 290–300 
4. Conclusions 
In this study, the MgO/MgAl 



Wavenumber (cm-1) 
spinel in biodiesel production reaction. Initially, MgAl 



was pre- pared with the combustion method by different types of fuel 
as  the  catalyst  support,  and  subsequently,  the  MgO  active  phase  was  loaded  on  the  support  surface  with  impregnation  method. 
Evaluat-  ing  the  synthesized  catalysts  by  various  fuel  types,  the  synthesized  sample  by  urea  showed the highest performance in 
the biodiesel production reaction with a conversion of 95.7%. The results of anal- yses presented that the as-prepared MgAl 

with  different  types  of  fuel  showed  various  structures,  morphologies,  surface 
areas and pore diameters which proved different performances of MgO/MgAl 





species  in  the  biodiesel  production.  As  described,  urea  was  selected  as  the  best  fuel  for  the  fabrication  of  the 
catalyst  support  due  to  its  high  crystallization  degree,  suitable  morphology  and  large  surface  area  and  pore  diameter  of  the 
synthesized  sup-  port.  Stability  test was performed on the optimum catalyst; after 6 replications, the catalyst activity decreased to 
about  83%  of  its  initial  activity  that  was  an  appropriate  value  for  the  synthesized  catalyst  (from  the  3rd  to  the  6th  rounds,  the 
activity remained almost constant). Finally, it can be said that the MgO/MgAl 

cat-  alyst  provides  good  potentials  for  industrialization  due  to 
its suit- able performance in biodiesel production as well as its simple and inexpensive preparation method. 


Catalyst loading = 3 (wt.%) 100 
Methanol/Oil = 12 molar 
95.7 Time = 3 h T=110°C 

MgO/MgAl2O4 
(G) 
Fig. 10. Influence of fuel ratio on catalytic performance of combustion synthesized MgO/MgAl 

84.5 
80 
60 
51.0 
58.0 
40 
20 
MgO/MgAl2O4 
MgO/MgAl2O4 
MgO/MgAl2O4 (CA) 
(EG) 
(U) Nanocatalyst 


nanocatalyst using various fuels. 
100 
95.1 
85.6 
84.1 83.2 82.9 

MgO/MgAl2O4 (FR=1.5) Catalyst loading = 3 (wt.%) Methanol/Oil = 12 molar 
Fig. 11. Activity of MgO/MgAl 

Time = 3 h 
100 
T=110°C 
80 
60 
40 
20 
1 2 3 4 5 6 
Run 


(U) nanocatalyst toward biodiesel production from sunflower oil. 
 
Acknowledgements 
The  authors  gratefully  acknowledge  Sahand  University  of  Tech-  nology  for  the  financial  support  of  the  research  as  well  as 
Iran Nan- otechnology Initiative Council for complementary financial supports. 
References 
[1] Zhou Q, Zhang H, Chang F, Li H, Pan H, Xue W, et al. Nano La 



as a heterogeneous catalyst for biodiesel synthesis by 
transesterification of Jatropha curcas L. oil. J Ind Eng Chem 2015;31:385–92. [2] Helwani Z, Aziz N, Bakar MZA, Mukhtar H, 
Kim J, Othman MR. Conversion of Jatropha curcas oil into biodiesel using re-crystallized hydrotalcite. Energy Convers Manage 
2013;73:128–34. [3] Shan R, Zhao C, Lv P, Yuan H, Yao J. Catalytic applications of calcium rich waste materials for biodiesel: 
current state and perspectives. Energy Convers Manage 2016;127:273–83. [4] Dharma S, Ong HC, Masjuki HH, Sebayang AH, 
Silitonga AS. An overview of engine durability and compatibility using biodiesel–bioethanol–diesel blends in 
compression-ignition engines. Energy Convers Manage 2016;128:66–81. [5] Zhou W, Konar SK, Boocock DGB. Ethyl esters 
from the single-phase base- 
catalyzed ethanolysis of vegetable oils. J Am Oil Chem Soc 2003;80:367–71. [6] Al-Jammal N, Al-Hamamre Z, Alnaief M. 
Manufacturing of zeolite based catalyst from zeolite tuft for biodiesel production from waste sunflower oil. Renewable Energy 
2016;93:449–59. [7] Liu H, Guo HS, Wang XJ, Jiang JZ, Lin H, Han S, et al. Mixed and ground KBr- impregnated calcined snail 
shell and kaolin as solid base catalysts for biodiesel production. Renewable Energy 2016;93:648–57. [8] Luque R. Algal biofuels: 
the eternal promise? Energy Environ Sci 
2010;3:254–7. [9] Torres-Rodríguez DA, Romero-Ibarra IC, Ibarra IA, Pfeiffer H. Biodiesel production from soybean and 
Jatropha oils using cesium impregnated sodium zirconate as a heterogeneous base catalyst. Renewable Energy 2016;93:323–31. 
[10] Yahya NY, Ngadi N, Jusoh M, Halim NAA. Characterization and parametric study of mesoporous calcium titanate catalyst 
for transesterification of waste cooking oil into biodiesel. Energy Convers Manage 2016;129:275–83. [11] Kwon EE, Yi H, Jeon 
YJ. Mechanistic investigation into water tolerance of non- 
catalytic biodiesel conversion. Appl Energy 2013;112:388–92. [12] Lotero E, Liu Y, Lopez DE, Suwannakarn K, Bruce DA, 
Goodwin JG. Synthesis of 
biodiesel via acid catalysis. Ind Eng Chem Res 2005;44:5353–63. [13] Sani YM, Daud WMAW, Abdul Aziz AR. Activity of 
solid acid catalysts for 
biodiesel production: a critical review. Appl Catal A: Gen 2014;470:140–61. [14] Sandesh S, Kristachar PKR, Manjunathan 
P, Halgeri AB, Shanbhag GV. Synthesis 
of biodiesel and acetins by transesterification reactions using novel CaSn(OH) 
6 heterogeneous base catalyst. Appl Catal A: Gen 
2016;523:1–11. [15] Xia S, Guo X, Mao D, Shi Z, Wu G, Lu G. Biodiesel synthesis over the CaO-ZrO 
2 solid base catalyst prepared by a urea-nitrate combustion 
method. RSC Adv 2014;4:51688–95. [16] Atadashi IM, Aroua MK, Abdul Aziz AR, Sulaiman NMN. The effects of catalysts 
in biodiesel production: a review. J Ind Eng Chem 2013;19:14–26. [17] Sánchez M, Avhad MR, Marchetti JM, Martínez M, 
Aracil J. Enhancement of the jojobyl alcohols and biodiesel production using a renewable catalyst in a pressurized reactor. 
Energy Convers Manage 2016;126:1047–53. [18] Avhad MR, Marchetti JM. A review on recent advancement in catalytic 
materials for biodiesel production. Renew Sustain Energy Rev 2015;50:696–718. [19] Mahdavi V, Monajemi A. Optimization of 
operational conditions for biodiesel 
production from cottonseed oil on CaO-MgO/Al 



solid base catalysts. J Taiwan Instit Chem Eng 2014;45:2286–92. [20] Umdu ES, 
Tuncer M, Seker E. Transesterification of nannochloropsis oculata 
microalga’s lipid to biodiesel on Al 



supported CaO and MgO catalysts. Bioresour Technol 2009;100:2828–31. [21] Wu H, Zhang 
J, Wei Q, Zheng J, Zhang J. Transesterification of soybean oil to biodiesel using zeolite supported CaO as strong base catalysts. 
Fuel Process Technol 2013;109:13–8. [22] Keane MA. Ceramics for catalysis. J Mater Sci 2003;38:4661–75. [23] Zhang P, Han 
Q, Fan M, Jiang P. Magnetic solid base catalyst CaO/CoFe 



for biodiesel production: influence of basicity and wettability of 
the catalyst in catalytic performance. Appl Surf Sci 2014;317:1125–30. [24] Mierczynski P, Chalupka KA, Maniukiewicz W, 
Kubicki J, Szynkowska MI, 
Maniecki TP. SrAl 



spinel phase as active phase of transesterification of rapeseed oil. Appl Catal B: Environ 2015;164:176–83. 
[25] Liu Q, Wang L, Wang C, Qu W, Tian Z, Ma H, et al. The effect of lanthanum doping on activity of Zn-Al spinel for 
transesterification. Appl Catal B: Environ 2013;136–137:210–7. [26] Sankaranarayanan TM, Shanthi RV, Thirunavukkarasu K, 
Pandurangan A, Sivasanker S. Catalytic properties of spinel-type mixed oxides in transesterification of vegetable oils. J Mol 
Catal A: Chem 2013;379:234–42. [27] Hashemzehi M, Saghatoleslami N, Nayebzadeh H. A study on the structure and 
catalytic performance of Zn 

Cu 
1Àx 
Al 



catalysts synthesized by the solution combustion method for the esterification reaction. C R 
Chim 2016;19:955–62. 
B. Rahmani Vahid, M. Haghighi / Energy Conversion and Management 134 (2017) 290–300 299 
[28] Foletto EL, Jahn SL, Muniz Moreira RdFP. Synthesis of high surface area 
MgAl 



nanopowder as adsorbent for leather dye removal. Sep Sci Technol 2009;44:2132–45. [29] Navaei Alvar E, Rezaei M, 
Navaei Alvar H, Feyzallahzadeh H, Yan Z-F. Synthesis 
of nanocrystalline Mgal 



spinel by using ethylene diamine as precipitation agent. Chem Eng Commun 2009;196:1417–24. [30] de 
Moraes Graziela Guzi, Oliveira APNd. Synthesis of the MgAl 



spinel obtained via combustion reaction using glycerine from the 
biodiesel as a fuel for producing cellular ceramics. Mater Sci Forum 2015;820:96–101. [31] Charghand M, Haghighi M, 
Aghamohammadi S. The beneficial use of ultrasound in synthesis of nanostructured Ce-doped SAPO-34 used in methanol 
conversion to light olefins. Ultrason Sonochem 2014;21:1827–38. [32] Parvas M, Haghighi M, Allahyari S. Degradation of 
phenol via wet-air oxidation over CuO/CeO2-ZrO2 nanocatalyst synthesized employing ultrasound energy: physicochemical 
characterization and catalytic performance. Environ Technol 2014;35:1140–9. [33] Estifaee P, Haghighi M, Babaluo AA, 
Rahemi N, Fallah Jafari M. The beneficial 
use of non-thermal plasma in synthesis of Ni/Al 



-MgO nanocatalyst used in hydrogen production from reforming of CH4/CO 

greenhouse gases. J Power Sources 2014;257:364–73. [34] Mathew CT, Vidya S, 
Koshy J, Solomon S, Thomas JK. Enhanced infrared 
transmittance properties in ultrafine MgAl 



nanoparticles synthesised by a single step combustion method, followed by hybrid 
microwave sintering. Infrared Phys Technol 2015;72:153–9. [35] Bai J, Liu J, Li C, Li G, Du Q. Mixture of fuels approach for 
solution combustion 
synthesis of nanoscale MgAl 



powders. Adv Powder Technol 2011;22:72–6. [36] Padmaraj O, Venkateswarlu M, Satyanarayana 
N. Structural, electrical and 
dielectric properties of spinel type MgAl 



nanocrystalline ceramic particles synthesized by the gel-combustion method. Ceram Int 2015;41:3178–85. [37] Rahmani Vahid 
B, Haghighi M. Urea-nitrate combustion synthesis of MgO/ 
MgAl 



nanocatalyst used in biodiesel production from sunflower oil: Influence of fuel ratio on catalytic properties and 
performance. Energy Convers Manage 2016;126:362–72. [38] Aliotta C, Liotta LF, La Parola V, Martorana A, Muccillo ENS, 
Muccillo R, et al. Ceria-based electrolytes prepared by solution combustion synthesis: The role of fuel on the materials 
properties. Appl Catal B: Environ 2016;197:14–22. [39] Madhusudhana HC, Shobhadevi SN, Nagabhushana BM, Chaluvaraju 
BV, Murugendrappa MV, Hari Krishna R, et al. Effect of fuels on conductivity, dielectric and humidity sensing properties of 
ZrO2 nanocrystals prepared by low temperature solution combustion method. J Asian Ceram Soc 2016;4:309–18. [40] Tarragó 
DP, Malfatti CdF, de Sousa VC. Influence of fuel on morphology of LSM powders obtained by solution combustion synthesis. 
Powder Technol 2015;269:481–7. [41] Scherrer P. Bestimmung der grösse und der inneren struktur von kolloidteilchen mittels 
röntgenstrahlen. Nachrichten von der Gesellschaft der Wissenschaften zu Gottingen 1918;26:98–100. [42] Granados ML, Poves 
MDZ, Alonso DM, Mariscal R, Galisteo FC, Moreno-Tost R, et al. Biodiesel from sunflower oil by using activated calcium 
oxide. Appl Catal B: Environ 2007;73:317–26. [43] Coenen JWE. Catalytic hydrogenation of fatty oils. Ind Eng Chem Fundam 
1986;25:43–52. [44] Abramoff MD, Magalhaes PJ, Ram SJ. Image processing with ImageJ. Biophoto 
Int 2004;11:36–42. [45] Yosefi L, Haghighi M, Allahyari S, Shokrani R, Ashkriz S. Abatement of toluene 
from polluted air over Mn/Clinoptilolite-CeO 

nanopowder: impregnation vs. ultrasound assisted synthesis with various Mn-loading. 
Adv Powder Technol 2015;26:602–11. [46] Sharifi M, Haghighi M, Abdollahifar M. Hydrogen production via reforming of 
biogas over nanostructured Ni/Y catalyst: effect of ultrasound irradiation and Ni-content on catalyst properties and performance. 
Mater Res Bull 2014;60:328–40. [47] Charghand M, Haghighi M, Saedy S, Aghamohammadi S. Efficient hydrothermal synthesis 
of nanostructured SAPO-34 using ultrasound energy: physicochemical characterization and catalytic performance toward 
methanol conversion to light olefins. Adv Powder Technol 2014;25:1728–36. [48] Rahmani F, Haghighi M, Mahboob S. CO 

-enhanced dehydrogenation of ethane over sonochemically synthesized Cr/clinoptilolite-ZrO 

nanocatalyst: Effects of ultrasound irradiation and ZrO 

loading on catalytic activity and stability. Ultrason Sonochem 2016;33:150–63. [49] Ahmadi F, 
Haghighi M, Ajamein H. Sonochemically coprecipitation synthesis of 
CuO/ZnO/ZrO2/Al 



nanocatalyst for fuel cell grade hydrogen production via steam methanol reforming. J Mol Catal A: Chem 
2016;421:196–208. [50] Khoshbin R, Haghighi M. Urea-nitrate combustion synthesis and physicochemical characterization of 
CuO-ZnO-Al2O3 nanoparticles over HZSM-5. Chin J Inorgan Chem 2012;28:1967–78. [51] Sajjadi SM, Haghighi M, Rahmani 
F. Dry reforming of greenhouse gases CH 

/ CO 

over MgO-promoted Ni-Co/Al 



-ZrO 

nanocatalyst: effect of MgO addition via sol-gel method on catalytic properties and 
hydrogen yield. J Sol-Gel Sci Technol 2014;70:111–24. [52] Saberi A, Golestani-Fard F, Sarpoolaky H, Willert-Porada M, 
Gerdes T, Simon R. Chemical synthesis of nanocrystalline magnesium aluminate spinel via nitrate-citrate combustion route. J 
Alloy Compd 2008;462:142–6. [53] Boroujerdnia M, Obeydavi A. Synthesis and characterization of NiO/MgAl 


4  nanocrystals  with  high  surface  area  by  modified  sol-gel 
method. Microporous Mesoporous Mater 2016;228:289–96. 
 
300 B. Rahmani Vahid, M. Haghighi / Energy Conversion and Management 134 (2017) 290–300 
[54] Li F-T, Zhao Y, Liu Y, Hao Y-J, Liu R-H, Zhao D-S. Solution combustion synthesis and visible light-induced photocatalytic 
activity of mixed amorphous and [55] crystalline C 
ˇ 
apek L, Kutálek MgAl 

P, O 
4 Smoláková nanopowders. L, Hájek Chem M, Eng Troppová J 2011;173:750–9. 
I, Kubicka D. The effect of thermal pre-treatment on structure, composition, basicity and catalytic activity of Mg/Al mixed 
oxides. Top Catal 2013;56:586–93. [56] Hájek M, Kutálek P, Smoláková L, Troppová I, C ˇ 
apek L, Kubicka D, et al. Transesterification of rapeseed oil by Mg-Al mixed oxides 
with various Mg/Al molar ratio. Chem Eng J 2015;263:160–7. [57] Brito A, Borges ME, Garín M, Hernández A. Biodiesel 
production from waste oil using MgÀAl layered double hydroxide catalysts. Energy Fuels 2009;23:2952–8. 
ˇ [58] Kutálek P, C 
apek L, Smoláková L, Kubicka D. Aspects of Mg-Al mixed oxide activity in transesterification of rapeseed oil in a 
fixed-bed reactor. Fuel Process Technol 2014;122:176–81. [59] Silva CCCM, Ribeiro NFP, Souza MMVM, Aranda DAG. 
Biodiesel production from soybean oil and methanol using hydrotalcites as catalyst. Fuel Process Technol 2010;91:205–10. [60] 
Barakos N, Pasias S, Papayannakos N. Transesterification of triglycerides in high and low quality oil feeds over an HT2 
hydrotalcite catalyst. Bioresour Technol 2008;99:5037–42. 

Você também pode gostar