Anda di halaman 1dari 17

Lihat diskusi, statistik, dan profil penulis untuk publikasi ini di: https://www.researchgate.

net/publication/321013619

Produksi-Tekanan-Drawdown Manajemen untuk Retak Horizontal Wells di Shale Gas-Formasi


(SPE-181.365-PA)

Artikel    di    SPE Reservoir Evaluasi & Engineering · November 2017

DOI: 10,2118 / 181.365-PA

CITATIONS Dibaca

5 207

4 penulis:

Matteo Marongiu-Porcu

Economides Konsultan
8 PUBLIKASI     14 CITATIONS    

30 PUBLIKASI     133 CITATIONS    


MELIHAT PROFIL

MELIHAT PROFIL

Hanyi Wang Philippe Enkababian

University of Texas di Austin Schlumberger Terbatas

43 PUBLIKASI     325 CITATIONS     27 PUBLIKASI     33 CITATIONS    

MELIHAT PROFIL MELIHAT PROFIL

Beberapa penulis publikasi ini juga bekerja pada proyek-proyek terkait:

Hidrolik Perekahan Modeling Dengan Metode Elemen Hingga Lihat proyek

P3D Modeling baru Hydraulic Patah Dakwah dan Penutupan Lihat proyek Ankit Mirani

Semua konten berikut halaman ini diunggah oleh Hanyi Wang pada Oktober 2018 05.

Pengguna telah meminta peningkatan file yang didownload.


Produksi-Tekanan-Drawdown Manajemen
untuk Retak Horizontal
Wells di Shale Gas-Formasi
Ankit Mirani 1, University of Houston; Matteo Marongiu-Porcu, Schlumberger;
Hanyi Wang, University of Texas di Austin; dan Philippe Enkababian, Schlumberger

Ringkasan
Perkembangan formasi shale-gas konvensional di Amerika Utara dengan sumur multifractured horisontal cukup matang untuk mengidentifikasi productionmalpractices dan penurunan
produktivitas yang abnormal umumnya diamati dalam 18-24months dari produksi awal. Tujuan utama dari penelitian ini adalah untuk mengatasi semua diketahui penyebab penurunan
produktivitas ini dan untuk mengembangkan sepenuhnya ditambah geomechanical / fl ow simulationmodel untuk mensimulasikan kondisi produksi tersebut.

Model ini meniru efek deplesi-induced in-situ variasi stres pada jangka pendek dan jangka panjang produktivitas dengan mempertimbangkan beberapa fenomena, seperti
matriks stres tergantung dan permeabilitas alam-patah serta penurunan konduktivitas hidrolik-fraktur disebabkan oleh menghancurkan proppantnya, deformasi, embedment,
dan patah-wajah merayap. evolusi matriks-permeabilitas, mengingat efek yang saling bertentangan con fl non-Darcy fl ow dan pemadatan, juga telah diperhitungkan dalam
model ini. solusi numerik untuk disederhanakan hidrolik-fraktur geometri planar kemudian diperoleh dengan menggunakan skema fi nite-elemen-metode.

Kasus sintetis didefinisikan untuk menyelidiki efek dari setiap fenomena individu pada jangka pendek dan produksi jangka panjang. Hasil penelitian
menunjukkan bahwa efek gabungan dari perubahan permeabilitas dalam matriks dan alami patah tulang serta kerugian konduktivitas di fraktur hidrolik
dapat mengakibatkan kerugian kumulatif-gas-produksi yang cukup besar. Model ini juga mereproduksi familiar fi tren lapangan-diamati, dengan produksi
jangka panjang lebih rendah sesuai dengan penarikan yang lebih tinggi. Perilaku ini dikaitkan dengan evolusi stres tergantung dari permeabilitas waduk
dan konduktivitas hidrolik-patah. Hasil penelitian menunjukkan bahwa mengabaikan efek dari salah satu hasil fenomena sebelumnya terlalu tinggi
pemulihan akhir. Selanjutnya,

Model ini sepenuhnya fleksibel dan memungkinkan pemodelan dan karakterisasi semua sangat berbeda (pada tingkat petrofisika) formasi shale-gas serta bahan
proppantnya digunakan untuk perawatan stimulasi. Model terpadu ini dapat digunakan untuk optimasi parameter kunci selama desain hidrolik-patah, untuk mendefinisikan
pencocokan sejarah produksi tuning, dan terutama sebagai alat prediksi untuk manajemen tekanan-penarikan.

pengantar
Kemajuan teknologi dan inovasi dalam rekah hidrolik selama dekade terakhir telah membuat shale gas yang ekonomis dan komoditas yang layak di
Amerika Utara. Namun, pengembangan dan produksi ini reservoir konvensional yang kompleks telah dikaitkan dif-kesulitan dan ketidakpastian terkait
dengan kinerja baik dan peramalan produksi. Awalnya, ukuran pemulihan ultimate perkiraan (EUR) dianggap memiliki hubungan langsung dengan tingkat
produksi awal. Hal ini telah mendorong operator untuk mengoperasikan sumur di penarikan tekanan yang sangat tinggi untuk melaporkan tingkat produksi
awal yang tinggi. Namun, ini biasanya disertai dengan penurunan produktivitas yang sangat-curam. Perilaku ini adalah pertama dirasionalisasikan dan
dijelaskan oleh Miller et al. (2010). Dalam studi mereka,

Selanjutnya, beberapa penulis dibawa ke cahaya perdebatan yang ada di industri mengenai apakah membatasi tingkat produksi atau menerapkan tekanan balik
(drawdown rendah) untuk sumur dapat menghasilkan EUR lebih tinggi dibandingkan dengan sumur produksi pada tingkat produksi terbatas. Beberapa penulis
mengembangkan model reservoir yang menggabungkan tekanan / stres fenomena-dependent dalam studi simulasi mereka untuk model perilaku lapangan ini dan
mengusulkan bahwa hasil penarikan yang lebih tinggi dalam penipisan lebih cepat dari reservoir, yang pada gilirannya menghasilkan tegangan jauh-tinggi yang efektif.
Hal ini menyebabkan cepat patah-konduktivitas dan matriks-permeabilitas kerugian, kemudian mengakibatkan kerugian produktivitas parah. Mereka juga melaporkan
bahwa praktek produksi (seperti pemilihan proppantnya, konsentrasi proppantnya,

Okouma Mangha et al. (2011) menyarankan penggunaan permeabilitas yang sangat-sederhana jangka eksponensial-pembusukan yang merupakan fungsi dari tegangan efektif bersih untuk
analisis dan pemodelan kasus penarikan yang berbeda. Mereka memperkenalkan “permeabilitas modulus” konsep, identik dalam bentuk dengan istilah kompresibilitas, untuk pemodelan variasi
sistem permeabilitas sebagai satu kesatuan.
Sejenisnya et al. (2011) disajikan studi banding dengan solusi analitis, semianalytical, dan numerik untuk pertandingan sejarah sejumlah besar data lapangan. solusi
numerik seperti menyumbang nonlinier dan penurunan produktivitas baik (misalnya, sifat tekanan-dependent, penurunan konduktivitas fraktur, dan Darcy non-faktor kulit).

perbaikan yang cukup dibuat oleh Eshkalak et al. (2014) dalam menjelaskan perilaku fracture- aliran dan matrix- fl ow potensial dengan menipisnya
cadangan tekanan. Pertama, penelitian ini dianggap model permeabilitas-pembusukan individu untuk matriks, hidrolik, dan patah tulang diinduksi. Kedua,
mereka yang tergabung efek gas-selip dan non-Darcy- fl ow mekanisme. Kedua perbaikan yang signifikan dibandingkan dengan model sebelumnya (Ilk et
al 2011;. Okouma Mangha et al 2011.). Namun, tidak ada upaya dilakukan untuk mengatasi perilaku permeabilitas alam-fraktur dengan perubahan
tegangan efektif bersih dan implikasi dari perubahan untuk kinerja waduk. Selain itu, permeabilitas patah tulang hidrolik dimodelkan sebagai peluruhan
eksponensial sederhana,

1 Saat ini dengan Essar Oil Ltd

hak cipta V C 2017 Society of Petroleum Engineers

Makalah ini (SPE 181.365) diterima untuk presentasi di Konferensi Tahunan Teknis SPE dan Pameran, Dubai, September 26-28 2016, dan direvisi untuk publikasi. Naskah asli diterima untuk meninjau 13 Jul 2016. Revisi Naskah
diterima untuk meninjau 27 Jun 2017. Kertas rekan disetujui 18 Juli 2017.

2017 SPE Reservoir Evaluasi & Engineering 1

ID: jaganm Waktu: 15:08 Saya Path: S: / REE # / Vol00000 / 170.031 / Comp / APPFile / SA-REE # 170.031
Britt et al. (2016) disajikan kajian komprehensif tentang perilaku alam-fraktur untuk kain batu yang berbeda dan dalam berbagai kondisi penarikan. Mereka menunjukkan
skenario produksi yang berbeda mungkin bagi beberapa formasi yang tidak konvensional aktif di mana penarikan hasil manajemen yang cukup produksi bene ts fi. Hasil
penelitian mereka menegaskan bahwa patah tulang alami memainkan peran penting dalam peningkatan produksi pada relatif “keras” formasi seperti Barnett sebagai lawan
yang relatif “lunak” formasi seperti Haynesville karena penutupan lengkap cepat mereka dalam kasus yang terakhir.

Efek proppantnya-diagenesis dan reaksi geokimia dengan in-situ fl UID menyajikan set kontroversial lain sumber potensial untuk penurunan produktivitas tergantung
waktu. Lee et al. (2009) dijelaskan diagenesis proppantnya sebagai tiga proses serial (pembubaran di butir / kontak biji-bijian, difusi pada antar muka air fi lm
memisahkan biji-bijian, dan curah hujan di dinding pori), tetapi meninggalkan belum terselesaikan ketidakpastian besar dalam kinetika reaksi yang terlibat. Menurut
Blauch et al. (2009), selama pasca-fraktur fl owback, yang fluida dapat berinteraksi dengan rock formation, berpotensi melarutkan garam larut; ion logam seperti besi,
barium, strontium, kalsium, dan magnesium; sulfat terlarut seperti gipsum, anhidrit, dan celestite; scale dan mikroba membentuk gas-; dan signifikan tingkat fi kan anion
termasuk karbonat, bikarbonat, sulfat, dan klorida. Blauch et al. (2009) maka hipotesis efek buruk dari potensi curah hujan dari salah satu bahan tersebut (skala
anorganik) dalam paket fraktur proppantnya, sedangkan Selamat et al. (2015) mengusulkan gagasan yang tidak didukung dan tidak-divalidasi bahwa pembubaran patah
tulang alami garam-disegel bisa memberikan manfaat efek resmi produktivitas baik dan EUR.

kemajuan substansial telah dibuat ke arah pemodelan dan simulasi cairan fl ow di waduk shale yang tidak konvensional, seperti model yang disajikan oleh Brown et al. (2009)
dan Yu dan Sepehrnoori 2014). Namun, industri ini masih kekurangan simulator reservoir yang kuat terintegrasi yang dapat mencakup semua mekanisme yang dikenal, seperti
non-Darcy fl ow, efek adsorpsi-layer, efek geomechanical untuk fl ow dalam matriks, perubahan proppantnya-properti (crushing, deformasi, embedment) , perubahan
patah-aperture / permeabilitas untuk aliran dalam patah tulang hidrolik, dan perubahan alam-patah-permeabilitas untuk aliran dalam patah tulang alami.

Pekerjaan yang disajikan dalam artikel ini dikandung untuk mengatasi aspek-aspek yang hilang dari pemodelan; kami mengembangkan simulator reservoir yang terintegrasi akuntansi untuk
berbagai tekanan / stres fenomena-dependent.
Kami dimasukkan efek non-Darcy fl ow, waduk pemadatan, dan adsorpsi lapisan pada evolusi matriks permeabilitas dengan deplesi. Kami menyelidiki dan model efek
menghancurkan proppantnya, deformasi proppantnya, dan embedment proppantnya dengan perubahan stres in-situ pada konduktivitas hidrolik-patah. Terakhir, kita menyumbang
variasi dalam permeabilitas alam-fraktur dengan perubahan deplesi-diinduksi stres in-situ. Sebuah numerical- aliran model simulasi mengintegrasikan fenomena ini kemudian
dikembangkan dan digunakan untuk merekomendasikan praktek produksi terbaik ditujukan terhadap peningkatan pemulihan jangka panjang.

Perlu disebutkan bahwa efek proppantnya-diagenesis dan reaksi geokimia dengan in-situ fluida belum dipertimbangkan dalam penelitian ini, diberikan pemahaman yang
buruk dan kurangnya model fundamental dalam literatur.

Model Formulasi
Reservoir Matrix. Arus dalam struktur nanometer-pori tidak dapat dijelaskan hanya oleh hukum Darcy. Klinkenberg (1941) identifikasi ed terjadinya selip gas dalam media
berpori dan mengamati bahwa gas yang sebenarnya aliran itu lebih dari yang diperkirakan dari hukum Darcy. Banyak penulis (Javadpour et al 2007;. Civan 2010;
Sakhaee-Pour dan Bryant 2012) Memiliki dikuanti fi kasi efek selip dengan memodifikasi faktor slip untuk menentukan permeabilitas matriks jelas sebagai fungsi dari jumlah
Knudsen. Wang dan Marongiu-Porcu (2015) mengembangkan sebuah fi Model ed uni untuk matriks permeabilitas dalam formasi shale-gas akuntansi untuk efek
Klinkenberg (1941) gas selip, pemadatan waduk, dan lapisan adsorpsi.

The fl uid- fl ow perilaku di nanopores melampaui rezim kontinum untuk mencapai slip, transisi, dan kondisi molekul bebas. Asumsi fl uid- fl ow mekanika klasik
(nol cairan kecepatan pada dinding pori) tidak memegang berlaku di luar continuum- aliran rezim. The Knudsen nomor, Kn, adalah parameter berdimensi yang
digunakan untuk membedakan rezim ow fl dalam saluran di mikro dan nano, dan didefinisikan sebagai rasio dari molekul berarti jalan bebas, k; dengan panjang
karakteristik saluran alir (dalam kasus kami, jari-jari pori r):

Kn ¼ k . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð1Þ
r:

Jumlah Knudsen di sebagian besar serpih adalah antara 10 3 dan 1 (Ziarani dan Aguilera 2012). Ini berarti bahwa slip- dan transition- fl ow rezim yang paling mungkin ditemui di
sebagian besar waduk shale-gas. The jelas matriks permeabilitas k Sebuah dapat direpresentasikan sebagai istilah umum berikut:

k Sebuah ¼ k 1 f ð Kn Þ; . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð2Þ

dimana k 1 adalah permeabilitas intrinsik untuk mengalir dari cairan yang ideal kental nonreacting dan f ð Kn Þ adalah istilah korelasi yang berhubungan permeabilitas matriks dan
permeabilitas intrinsik. Sakhaee-Pour dan Bryant (2012) mengembangkan korelasi, diverifikasi oleh data eksperimen, yang berkaitan istilah korelasi, f ð Kn Þ, dengan jumlah Knudsen
untuk slip- dan transition- fl ow rezim:

1 þ 5 Kn Menyelinap rezim
f ð Kn Þ ¼ Transisi rezim: . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð3Þ
0: 8453 þ 5: 4567 Kn þ 0: 1633 Kn 2

Jumlah Knudsen berkaitan dengan tekanan pori dan jalur bebas rata-rata dari molekul gas (Civan et al. 2011). Jadi, dengan perubahan tekanan reservoir, yang jelas
matriks permeabilitas juga akan berubah. Ini telah digunakan untuk model non-Darcy- efek fl ow ow ke dalam model fl kami. Sebuah diskusi rinci diberikan oleh Wang dan
Marongiu-Porcu (2015).
Berikutnya, kami menggabungkan efek pemadatan waduk (efek geomechanical) ke dalam evolusi matriks-permeabilitas. Telah ditemukan (Bes Kok dan
Karniadakis 1999; Aguilera 2002) bahwa intrinsik matriks permeabilitas kira-kira sebanding dengan kuadrat jari-jari pori:

k 1 / r 2: . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð4Þ

Radius pori akan berkurang dengan menurunnya tekanan pori karena peningkatan tegangan efektif. Oleh karena itu, jelas matriks permeabilitas juga akan
menurun karena efek geomechanical.
Mekanisme gas penyimpanan dalam reservoir shale gas secara konseptual berbeda dari reservoir konvensional. gas alam teradsorpsi pada bahan organik hadir dalam
serpih dan, dalam beberapa kasus, mineral tanah liat tertentu. Ini lapisan teradsorpsi-gas mengkonstriksi lanjut

2 2017 SPE Reservoir Evaluasi & Engineering

ID: jaganm Waktu: 15:08 Saya Path: S: / REE # / Vol00000 / 170.031 / Comp / APPFile / SA-REE # 170.031
daerah untuk fl ow melalui struktur pori, sehingga mengurangi radius pori-pori yang efektif. Sebagai tekanan dalam pori-pori mengurangi, gas diserap dan efektif radius pori
meningkat, mengakibatkan peningkatan dalam matriks permeabilitas jelas. Ia telah mengemukakan (Valko' dan Lee 2010) yang desorpsi metana mungkin ikut bertanggung
jawab atas relatif panjang dan fl di ekor produksi yang telah diamati di beberapa waduk shale. Langmuir isoterm adalah gas-adsorpsi model / desorpsi yang paling sering
digunakan untuk karakterisasi reservoir shale-gas. Xiong et al. (2012) mengusulkan model interpolasi ketebalan lapisan gas-adsorpsi dengan menggunakan hubungan
fungsional Langmuirtype, yang digunakan oleh Wang dan Marongiu-Porcu (2015) untuk memperoleh korelasi menggabungkan efek dari lapisan adsorpsi pada radius
pori-pori yang bervariasi dan permeabilitas intrinsik dikutip sebelumnya. Ini diberikan sebagai

PP
rm 0: 5 C /
L
r ¼ r0 dm ; . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð5Þ
0: 1
1þP
PL
2664 3775 2
PP
rm 0: 5 C /
k10r0 dm L
0: 1
1þP
PL
k1¼ ; . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð6Þ
r20

dimana r 0 dan k 1 0 yang jari-jari pori dan permeabilitas intrinsik diukur pada tekanan atmosfer P 0; C / adalah batu bahan konstan;
d m adalah diameter molekul gas; P adalah reservoir (pori) tekanan; dan P L adalah tekanan Langmuir.
The jelas matriks permeabilitas adalah fungsi dari interaksi dari tiga mekanisme: Jika efek geomechanical mendominasi, matriks permeabilitas akan menurun dengan
meningkatnya deplesi, sedangkan jika desorpsi dan non-Darcy mekanisme mendominasi, matriks permeabilitas akan meningkat dengan meningkatnya deplesi. Metodologi
diadopsi untuk pengembangan korelasi ini adalah di luar lingkup artikel ini, dan pengguna yang tertarik disebut Wang dan Marongiu-Porcu (2015), dari mana kita
mereproduksi Gambar. 1. Gambar. 1 menggambarkan ini interplays matriks-permeabilitas kompleks.
Average Apparent Permeability

160 170
in SRV (nd)

140 150

120 130 Non-Darcy flow Non-Darcy flow, geomechanics Non-Darcy


flow, geomechanics, adsorption layer
100 110

0 5 10 Waktu Produksi
15 (tahun)
20 25 30

Fig. 1—Average apparent-matrix-permeability variations showing effects of different mechanisms/models, fromWang and MarongiuPorcu (2015) (their Fig. 7). SRV 5 stimulated reservoir volume.

Pers. 5 dan 6 dapat sekarang digunakan dalam persamaan ow fl berhasil menggabungkan efek geomechanical dan efek adsorpsi-layer, bersama dengan non-Darcy- fl
ow efek seperti yang dijelaskan sebelumnya, pada perubahan permeabilitas matriks.
Di bawah asumsi dari media berpori elastis sempurna, sehingga tidak ada deformasi plastik terjadi, persamaan konstitutif berikut dapat dinyatakan dalam
tegangan efektif ( r aku j;), regangan ( e aku j), dan tekanan pori ( P):

r aku j ¼ 2 G e aku j þ 2 Gv . .j; . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð 7 Þ


Sebuah P d saya;
1 2 v e kk d saya; j

dimana G adalah modulus geser; v adalah rasio Poisson; e kk merupakan strain volumetrik; d saya; j adalah Kronecker d didefinisikan sebagai kesatuan untuk
saya ¼ j dan sebagai 0 untuk saya 6¼ j; dan Sebuah adalah Biot ini efektif-stres koefisien, yang diasumsikan kesatuan dalam penelitian ini. Hubungan strain-perpindahan dan persamaan keseimbangan
yang didefinisikan sebagai

e aku j ¼ 1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð8Þ
2 ð u saya; j þ u Ji Þ;

r aku j; j þ F saya ¼ 0; . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð9Þ

dimana u saya dan F saya adalah komponen perpindahan dan kekuatan tubuh bersih di saya- arah, masing-masing. Menggabungkan pers. 7 sampai 9, kita memiliki persamaan Navier
dimodifikasi dalam hal perpindahan bawah kombinasi diterapkan-stres dan variasi pori-tekanan:

G r u saya þ G Sebuah P d saya; j þ F saya ¼ 0:. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð 10 Þ


1 2 vu j; ii

Gambar. 1, direproduksi fromWang dan Marongiu-Porcu (2015), menunjukkan efek dari mekanisme yang berbeda pada evolusi permeabilitas. Jelas, peramalan
shale-gas kinerja jangka panjang akan sangat terpengaruh jika efek ini tidak benar dipertimbangkan.

Fraktur hidrolik. fraktur hidrolik umumnya tetap terbuka dan konduktif dengan penggunaan bahan proppantnya; setelah terpapar penipu fi ning stres untuk jangka memanjang
waktu, menghancurkan proppantnya dan proppantnya deformasi penyebab degradasi proppantnya ini

2017 SPE Reservoir Evaluation & Engineering 3

ID: jaganm Time: 15:08 I Path: S:/REE#/Vol00000/170031/Comp/APPFile/SA-REE#170031


mechanical and physical properties. Furthermore, this leads to reductions in fracture width and fracture-pack permeability because of intermixing and plugging.

In addition, the interaction between the proppant and fracture surfaces under the same confining (closure) stress can result in embedment (plastic indentation and
subsidence of the proppant into the softer fracture walls), decreasing the fracture aperture and impairing conductivity. Fracture conductivity C f is expressed as a product of
the fracture permeability k f and the width w of the fracture:

C f ¼ k f w: . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð 11 Þ

Experimental studies on modeling fracture conductivity have been reported by many researchers (Lehman et al. 1999; Barree et al. 2003; McDaniel et al. 2010).
There have been a few analytical models developed for modeling proppant embedment, proppant crushing, and proppant deformation (Chin et al. 2000; Brannon et al.
2004; Osholake et al. 2012; Eshkalak et al. 2014). However, there have not been relevant attempts to develop a comprehensive fracture-conductivity model that can be
implemented directly into a flow simulation model to capture pressure/stress-dependent changes in fracture conductivity during the production life of shale-gas wells.

In this study, we developed an analytical work flow to model fracture-conductivity variations with changing effective stress. Daneshy (2004) proposed the
“off-balance” fracture-propagation model describing the presence of multiple branches and small shear fractures caused by the presence of planes of weakness
within the rock. Fig. 2 explains the conceptual difference between the classic planar-fracture geometry, the “off-balance” fracture model, and the complex-fracture
model. The work proposed in this study considers a fracture model derived from the following assumptions:

• A planar primary hydraulic fracture has been considered along with the presence of secondary fractures, as described by the “offbalance” fracture model and
shown in Fig. 3.
• The primary hydraulic fracture has been considered fully propped, and the proppant distribution is considered to be full monolayer (Brannon et al. 2004), as shown
in Fig. 4.
• The secondary fractures have been considered unpropped for simplicity, and their behavior has been modeled in a way that is similar to that for the natural
fractures, which is discussed later in this text.

σ min

Simple Off-balance Complex network

Fig. 2—Basic fracture-growth models, from Daneshy (2004) (his Fig. 1).

Horizontal well

Complex-naturalfracture
system

Primary hydraulic fracture

Reservoir matrix

Single Biwing
fracture unit

Fig. 3—Schematic of the composite fracture model representing a unit pattern showing the primary hydraulic fracture and complex-natural-fracture system.

4 2017 SPE Reservoir Evaluation & Engineering

ID: jaganm Time: 15:08 I Path: S:/REE#/Vol00000/170031/Comp/APPFile/SA-REE#170031


Fig. 4—Schematic of a planar fracture containing full-monolayer proppant distribution, from Brannon et al. (2004) (their Fig. 3).

According to Eq. 11, the fracture conductivity can change either because of changes in the fracture (proppant-pack) permeability or because of changes in the
fracture width. Through this study, we have modeled changes in both of these parameters and developed an analytical solution for fracture-conductivity variations as a
function of in-situ effective stress.
Proppant Crushing: Modeling Changes in Fracture-Proppant-Pack Permeability. Data describing the stress-dependent evolution in proppant-pack conductivity are
among the most common and easy-to-find data in the technical literature, given the existence of specific procedures for measuring the properties of proppants used in
hydraulic fracturing defined by joint committees of the International Organization for Standardization (ISO) and the American Petroleum Institute (API) members
(Kaufman et al. 2007).
The codified procedures for testing of proppants followed by roughly 12 facilities worldwide are the now-obsolete API RP 61
(1989) and the trustable standard ISO 13503-5 ( 2006).
Although these two recent procedures have enabled users to evaluate and to compare proppant characteristics under the specifically described test conditions for
use in hydraulic-fracturing operations (Kaufman et al. 2007), their limitation resides in the necessary conversion to infer the values of proppant-pack permeability (from
the provided proppant-pack conductivity) that are then scaled up into discrete-grid models that constitute the input for any 2D or full 3D numerical reservoir simulator.
Even within the most-advanced autogridding algorithms (Cipolla et al. 2011a) applied to generate the unstructured grid for the potentially severe complex-fracture
patterns predicted by the most-advanced complex-fracture models (Weng et al. 2014), local grid refinement rarely determines cells with characteristic dimensions less
than 1 ft, whereas reported fracture widths in such unconventional reservoirs are on the order of millimeters.

The newly developed integrated reservoir simulator presented in this article accounts for crushing-induced evolution of proppantpack permeability directly, by use
of the Berg (1970) empirical relationship that links morphological variables such as grain size, shape, and sorting to the permeability of the fracture proppant pack k f:

k f ¼ 5:1 10 6/5 MPD 2 e 1:385 ð MPD P 10 Þ 1; 000; . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð 12 Þ

where / is the pack porosity, MPD is the mean particle diameter, and P 10 is the particle size corresponding to the P10 value on the cumulative proppant-size
distribution. Eq. 12 is an empirical relationship for varying fracture permeability with the morphological proppant variables; our aim is to obtain a relationship between
the fracture permeability and effective stress. Thus, if we can relate variation of the petrophysical parameters (MPD, P10, and /) with changing effective stress, we can
then obtain a relationship between fracture permeability and effective stress, as conceptually shown in Fig. 5.

Known Unknown Effective


Fracture
MPD/P10/ φ stress/confining
permeability
pressure

Fig. 5—Representation of the known correlations and a flow chart for determining the relationship between fracture permeability and confining pressure.

We propose the use of sieve-size analysis to determine MPD and P10 proppant size as a function of the confining effective stress. Mesh analysis involves
exposing a proppant sample to different confining stresses and subsequently passing it through a series of meshes to generate a probability size distribution that can
provide the MPD (P50 value) and the P10 proppant size.
We start from the results of a sieve-size analysis of a 20/40 natural sandpack exposed to different confining-stress conditions presented by Schubarth and
Milton-Tayler (2004). These are organized in Table 1.
Using the data in Table 1, we then generate a cumulative probability-distribution chart for the proppant pack corresponding to each sieve size, presented in Table 2 and
Fig. 6. This kind of data can be used to determine MPD and P10 sizes under the range of investigated confining stresses and to study the pressure/stress-dependent
fracture-pack-permeability deterioration, as in Eq. 12.
This mimics actual in-situ mechanisms, where during depletion the effective stress increases, resulting in changes in the values of MPD and P10 size because of the
proppant gradually crushing.
Proppant Embedment, Deformation, and Creep: Modeling Changes in Fracture Width. Proppant embedment has been previously defined as a stress-driven plastic
indentation and subsidence of the proppant into the softer fracture walls, decreasing the fracture aperture and impairing conductivity.

Chang and Zoback (2008) and Sone and Zoback (2010) have furthermore addressed the existence of a prefailure proppant elastic deformation (directly related to the
magnitude of the confining stress) and a viscoelastic time-dependent deformation, defined as creep.

2017 SPE Reservoir Evaluation & Engineering 5

ID: jaganm Time: 15:08 I Path: S:/REE#/Vol00000/170031/Comp/APPFile/SA-REE#170031


Mesh Size Unstressed 1,000 psi 2,000 psi 3,000 psi 4,000 psi 5,000 psi 6,000 psi 7,000 psi

18 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00

20 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00

25 0.03 0.03 0.03 0.03 0.03 0.03 0.03 0.02

30 0.17 0.15 0.15 0.15 0.15 0.14 0.12 0.10

35 0.45 0.44 0.44 0.44 0.43 0.40 0.31 0.27

40 0.27 0.28 0.28 0.27 0.27 0.25 0.21 0.20

45 0.65 0.08 0.08 0.08 0.08 0.09 0.10 0.11

50 0.02 0.02 0.02 0.02 0..02 0.03 0.04 0.05

60 0.01 0.01 0.01 0.01 0.02 0.03 0.06 0.07

70 0.00 0.00 0.00 0.00 0.00 0.01 0.04 0.04

80 0.00 0.00 0.00 0.00 0.01 0.03 0.11 0.15

Table 1 — Sieve-size distribution (in fraction) for a 20/40 proppant-pack sample under different confining stresses. Data are extracted from Schubarth and Milton-Tayler (2004).

Mesh Size Unstressed 1,000 psi 2,000 psi 3,000 psi 4,000 psi 5,000 psi 6,000 psi 7,000 psi

18 1.00 1.00 1.00 1.00 1.00 1.00 1.00 1.00

20 1.00 1.00 1.00 1.00 1.00 1.00 1.00 1.00

25 0.97 0.97 0.97 0.97 0.97 0.97 0.98 0.98

30 0.81 0.82 0.82 0.82 0.82 0.83 0.86 0.88

35 0.36 0.38 0.38 0.38 0.39 0.43 0.55 0.61

40 0.09 0.10 0.10 0.11 0.12 0.18 0.35 0.41

45 0.03 0.03 0.03 0.03 0.04 0.10 0.25 0.31

50 0.01 0.01 0.01 0.01 0.02 0.07 0.21 0.26

60 0.00 0.00 0.00 0.00 0.01 0.04 0.15 0.19

70 0.00 0.00 0.00 0.00 0.01 0.03 0.11 0.15

80 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00

Table 2 — Inverse cumulative probability distribution of proppant size under different confining stresses.

0 psi 1,000 psi


0.9 1

0.8 2,000 psi 3,000 psi

0.7 4,000 psi 5,000 psi

0.6
6,000 psi 7,000 psi
Probability Distribution

0.5

0.4

0.3
MPD
0.2

0.1

0
0 P10 10 20 30 40 50 60 70 80 90

Mesh Size

Fig. 6—Cumulative probability distribution of proppant size under different confining stresses.

Guo and Liu (2012) developed analytical models to calculate the proppant embedment and proppant deformation under a net confining pressure p c. They revealed a
linear dependence of elastic embedment and elastic deformation against the confining stress, as shown in Eqs. 13 and 14, respectively:

MPD 1
D PE ¼ 2 p c 1 v 2 2
; . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð 13 Þ
2 E1

D PD ¼ 1:04 ð MPD Þ p c ð 1 v 21 Þ ; . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð 14 Þ
E1

where v i is Poisson’s ratio of the material and E i is Young’s modulus. Subscripts 1 and 2 refer to the proppant and matrix, respectively.
The time-dependent creep deformation is modeled by use of the Kelvin viscoelastic model, as presented in the Shi et al. (2009) adaptation of the classic model:

6 2017 SPE Reservoir Evaluation & Engineering

ID: jaganm Time: 15:08 I Path: S:/REE#/Vol00000/170031/Comp/APPFile/SA-REE#170031


( " #)
t

e ð t Þ ¼ MPD 1 þ ð 1 2 v 22 Þ 2 pc1 e s ; . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð 15 Þ
2 G1 3

where G 1 is the shear modulus of the proppant material. The subscript convention remains the same as described previously.
The total net reduction in fracture width is then given as the summation of the proppant embedment, proppant deformation, and the time-dependent creep
deformation:

D w ¼ D PE þ D PD þ e ð t Þ: . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð 16 Þ

A hydraulic-fracture-conductivity multiplier can then be introduced, defined as a function of the fracture conductivity at any confining stress C fp c and the unaltered
fracture conductivity C f 0:

CM ¼ C fp c ; . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð 17 Þ
Cf0

where CM is conductivity multiplier. This hydraulic-fracture-conductivity multiplier is conveniently used in the integrated reservoir simulator presented in this article to
capture pressure/stress-dependent changes concurrently in the fracture-proppant-pack permeability (work flow in Fig. 5 and Eq. 12) and in the fracture-proppant-pack
width (Eq. 16) during the production life of shale-gas wells.

Natural Fractures. Natural fractures exist in many oil and gas formations and are the principal source of flow capacity in many of these reservoirs. Changes in
fracture-flow capacity and conductivity are expected to have an important influence on reservoir performance (Palmer et al. 2007; Vega Navarro 2012; Cho et al.
2013).
In principle, the spatial distribution of natural fractures can be modeled as a discrete-fracture network (DFN). The properties of natural fractures are estimated by
use of “ant tracking” of seismic attributes along with regional fracture trends from image logs, following the procedure presented by Offenberger et al. (2013). In
recent years, applications of the seismic attribute (Wilson et al. 2014) and its azimuthal inversion (Far et al. 2013) have been performed as novel work flows to model
more-accurate DFNs and to understand natural-fracture variability in unconventional reservoirs. Thachaparambil (2015) introduced a new method for extracting 3D
DFNs from seismic attributes, in correlation with well data.

Combining hydraulic-fracture modeling (Weng et al. 2014) and microseismic data (Shoemaker et al. 2015; Zakhour et al. 2015a; Ajisafe et al. 2016), the modeled
DFN can then be finely calibrated. Furthermore, the properties of the propped fracture and unpropped fracture can be determined through conductivity testing under
various pressures and confining stresses (Ghassemi and Suarez-Rivera
2012) so that the parameters defining transport capacity of fractures can be determined.
One of the major tenets in shale-gas (as well as shale-oil) stimulation is the beneficial activation of the contacted natural-fracture network with fracturing fluid,
caused by tensile or (more often) shear failure. Such poorly sealed natural fractures are generally reported to interact heavily with the hydraulic fractures during the
injection treatments, serving as preferential paths for the growth of a complex-fracture network. The reactivation of such pre-existing planes of weakness (i.e., natural
fractures, microfractures, and fissures) is well-documented and observed in microseismic monitoring (Cipolla et al. 2011b, c; Peyret et al. 2012; Zakhour et al. 2015b).

However, these natural fractures are only poorly propped or completely unpropped, given the manifested difficulty in “turning the corners” encountered by the
proppant transported by fracturing fluids. Therefore, the activated natural-fracture-network conductivity is susceptible to severe pressure/stress-dependent conductivity
impairment during depletion and must be taken into consideration in our integrated reservoir simulator. Liu et al. (2013) modeled the stress-dependent behavior of
fractures and cracks in porous media by use of the framework of Hooke’s law. Their work conceptualizes the rock matrix as having two parts: a “soft” part that obeys a
naturalstrain-based Hooke’s law, and a “hard” part that follows the engineering-strain-based Hooke’s law (Liu et al. 2013). Their study documents that the fracture
aperture of an unpropped natural fracture is a sum of two functions: one linearly varies with the effective stress and the other varies nonlinearly with change in effective
stress. Liu et al. (2013) provide a very descriptive macroscopic model, but their study does not rely on the detailed description of the small-scale structures (namely,
fracture-roughness peaks, as in the case of natural fractures) to describe the behavior of fractures. Although several stress-dependent-permeability correlations for
natural fractures have been discussed in the literature (Goodman 1974; Barton et al. 1985), many of them rely on the estimation of empirical parameters that serve as
curve-fitting tools (Cho et al. 2013).

In this article, we propose the use of the Raghavan and Chin (2002) model, which considers an exponential decline in the naturalfracture permeability:

k fn ¼ k fni e d f p c ; . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð 18 Þ

where k fni is the initial natural-fracture permeability and d f is the compressibility of the rock matrix. This correlation given by Eq. 18 is simple to use and does not require
extensive data, which is the reason for its widespread applicability (Cho et al. 2013).
Laboratory measurement of natural-fissure permeability as a function of effective stress performed by Britt et al. (2016) indicates, however, that the application of
an exponential decay of natural-fracture permeability may be too simplistic. Their experiments, performed with samples from “hard,” “medium,” and “soft” rock types,
clearly show that the evolution of the natural-fissure permeability under high effective stress is a function of the rock mechanical properties. Hard rock maintains a
much-higher natural-fissure permeability (two orders of magnitude higher) than soft rock, which deforms plastically upon closure and seals the fissure. Thus, the
application of an exponential decay of natural-fracture permeability, as used in our study, must be dependent on a detailed understanding of the rock matrix fabric
under consideration. The use of Eq. 18 may result in overestimation of cumulative production of gas at late to very-late times of the producing life of the well in soft
formations. This is because the exponential-decline nature of Eq. 18 will not be able to account for complete fracture closure, and some amount of permeability will
always be retained in the natural fractures.

In theory, such much-more-severe conductivity reduction in soft rock could be modeled by accounting for stress-dependent aperture deterioration of the contacted
natural-fracture network with fracturing fluids. Under exposure to an increasing net confining stress, and because the rock roughness peaks with the fracture surface, the
peaks of one rock face embed into the opposite face or deform, which results in decreasing the natural-fracture-network local aperture and, in turn, fracture conductivity.
Roughness peak embedment and deformation are serious issues in soft shale formations because of their low elastic moduli.

2017 SPE Reservoir Evaluation & Engineering 7

ID: jaganm Time: 15:08 I Path: S:/REE#/Vol00000/170031/Comp/APPFile/SA-REE#170031


Recasting the Guo and Liu (2012) analytical model (i.e., substituting the propped with the natural-fracture roughness peaks made of the same material as the
formation), the linear dependence of elastic embedment and elastic deformation against the confining stress could be simplified as shown in Eqs. 19 and 20,
respectively:

1
D PE ¼ 2 p c 1 v 2 R ed . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð 19 Þ
2 E;

D PD ¼ 1:04 ð R ed Þ p c ð 1 v 2 Þ ; . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð 20 Þ
E

where v is Poisson’s ratio, E is the Young’s modulus of the formation, and R ed is the roughness equivalent diameter representing the equivalent MPD of the roughness of
one natural-fracture face in contact with the opposite fracture face if such roughness was a proppant particle.

The total net reduction in natural-fracture aperture is then given as the summation of the roughness embedment and deformation:

D w ¼ D PE þ D PD ¼ 2:04 ð R ed Þ p c ð 1 v 2 Þ : . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð 21 Þ
E

The roughness equivalent diameter R ed could be determined empirically and included into Eq. 21 to obtain a formulation of naturalfracture-aperture reduction
deriving from formation mechanical properties. Such an approach, integrating the formation mechanical properties in the natural-fracture-conductivity evolution as
effective stress increases, would better represent the soft-rock cases (significant drop in aperture) compared with hard-rock cases where aperture reduction will be less
dramatic.
Another aspect of modeling flow in natural fractures is the permeability enhancement of natural fractures during fracturing treatment. As suggested by Zhang
and Li (2016), fracture shearing is a very important and irreversible contribution to increased fracture-surfaces roughness after fracture closure. However, for the
purposes of the present long-term production model, the initial naturalfracture permeability k fni used in Eq. 18 can be properly tuned to account for any initial
condition reflecting shearing or any other envisioned permeability enhancement of natural fractures during the hydraulic-fracturing treatments.

Model Setup
So far, we addressed all supported known pressure/stress-dependent phenomena that have an effect on the productivity decline of shalegas wells, which we now
incorporate into our integrated reservoir simulator; all the derived coupled equations are solved numerically, by use of the previous numerical framework fromWang and
Marongiu-Porcu (2015).
Table 3 shows all the input parameters used for the presented analysis, which include reservoir conditions, drainage geometry, hydraulic-fracture conductivity,
real-gas properties, and gas-adsorption parameters.
A realistic synthetic DFN was modeled by use of a standard deviation equal to 10 to mimic variability of two orthogonal subsets (north/south and east/west), with
fracture intensity (defined as total fracture area/fracture control volume) equal to 0.05. DFN length was modeled between 10 and 35m by use of the following
power-law distribution:

Count ¼ 100 length 1:25: . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð 22 Þ

Natural-fracture initial permeability and aperture are also listed in Table 3.


Although DFN reactivation in shear mode could have been implemented, with consequent further degradation of natural-fracture permeability, at this stage we
prefer not to introduce any hydromechanical coupling and to approach the flow simulation as a singleporosity problem. We use a finite-element-analysis scheme
(Zienkiewicz and Taylor 2005) derived from the Newton-Raphson method (Ypma 1995), which requires evaluating a Jacobian matrix starting from initial-guess values
to approximate the solution throughout successive iterations.

All the variables in the system are updated at the end of each time increment and entered as initial values at the start of the next time increment.

Although it is possible to simulate an entire section of a horizontal well with multiple fractures, it is more efficient to simulate a unit-drainage pattern for a
single-wing planar-fracture model with the induced secondary fractures in the “off-balance” mode (as conceptually illustrated in Fig. 2) and apply symmetric
conditions along the boundaries.
As in Wang and Marongiu-Porcu (2015), to make a consistent assessment on how the non-Darcy flow, adsorption layer, and geomechanical effects affect the
shale-matrix permeability during production, the reference input values for porosity, intrinsic permeability, and pore radius for all models are considered at standard
conditions under laboratory environment, whereas their corresponding values at initial reservoir conditions are calculated by use of in-situ
reservoir-pressure/temperature and reservoir-stress conditions. The rock material constant C/ ( which appears in Eqs. 5 and 6), is set to be 0.035, which is the
median value of laboratory measurements (Dong et al. 2010); the diameter of adsorption-gas molecules ( d m) is assumed to be 0.414 nm, which is the diameter of a
methane molecule.

The propped-hydraulic-fracture sections are modeled by use of the same sieve-size analysis of a 20/40 natural sandpack exposed to different confining-stress
conditions (Schubarth and Milton-Tayler 2004) introduced in Tables 1 and 2 and Fig. 6. The pressure/stressdependent fracture-pack-permeability deterioration is
modeled implementing Eq. 12 in our numerical framework.

Results and Discussion


Effect of Pressure-Dependent Phenomena on Long-Term Well Performance. To demonstrate the pressure/stress-dependent changes in matrix permeability and
hydraulic- and natural-fracture conductivity, we consider the model setting described previously, and we compare two 10-year-production simulations with and
without the incorporation of the modeled pressure/stress-dependent phenomena. Fig. 7 shows the results of these two scenarios. We can observe a difference of
approximately 30% in the cumulative production after the simulated 10 years, reflecting the deleterious effect of the depletion-related increase of the effective stress
upon matrixand fracture-flow potential.

8 2017 SPE Reservoir Evaluation & Engineering

ID: jaganm Time: 15:08 I Path: S:/REE#/Vol00000/170031/Comp/APPFile/SA-REE#170031


Input Parameters Value

Initial reservoir pressure p i 34.447 MPa

Wellbore pressure p wf 3.45–13.79 MPa

Intrinsic permeability measured at laboratory conditions k ∞0 100 nd

Hydraulic-fracture initial permeability k fi 10 4 md

Natural-fracture initial permeability k nfi 10 3 md

Formation porosity measured at laboratory conditions φ 0 0.08

Pore radius measured at laboratory conditions r 0 5 nm

Material constant for porosity change C φ 0.035

Hydraulic-fracture proppant-pack porosity φ f 0.35

Hydraulic-fracture width d f 0.001 m

Hydraulic-fracture proppant-pack areal concentration C p 4 lbm/ft 2

Natural-fracture aperture a f 0.0001 m

Fracturing spacing Y e 55 m

Reservoir thickness H 100 m

Drainage length parallel to the hydraulic fracture X e 122 m

Hydraulic-fracture half-length x f 76.2 m

Total number of drainage units n 20

Reservoir temperature T 366 K

Density of formation rock ρ m 2600 kg/m 3

Langmuir volume constant V L 0.0113 m 3/ kg

Langmuir pressure constant P L 10.34 MPa

Diameter of adsorption-gas molecules d m 0.414 nm

Average molecular weight M 16.04 g/mol

Critical temperature of mix gas T c 191 K

Critical pressure of mix gas P c 4.64 MPa

Horizontal minimum stress S h 37.92 MPa

Horizontal maximum stress S H 41.37 MPa

Overburden stress S v 44.82 MPa

Biot’s constant α 1

Young’s modulus E 25 GPa

Poisson’s ratio v 0.22

Table 3 — Input parameters for base-case simulation. REE181365 DOI:


Cumulative Gas Production (MMcf)

678

345 No pressure-dependent phenomena, 500 psi

012 Pressure-dependent phenomena, 500 psi

0 500 1,000 1,500 2,000 2,500 3,000 3,500 4,000

Time (days)

Fig. 7—Comparison of cumulative gas production with and without the incorporation of pressure/stress-dependent phenomena.

2017 SPE Reservoir Evaluation & Engineering 9

ID: jaganm Time: 15:09 I Path: S:/REE#/Vol00000/170031/Comp/APPFile/SA-REE#170031


Fig. 8 shows the decoupled effects of matrix, hydraulic-fracture, and natural-fracture pressure/stress-dependent phenomena on the cumulative production plotted in
Fig. 7. The four cases considered in the analysis are shown in Table 4. Specifically, Case 2 shows that incorporating the pressure/stress-dependent permeability
changes in only the matrix leads to higher cumulative production during the early life of the well, but eventually the production rate declines much faster, leading to a
lower overall recovery. This result follows what is discussed previously in this text and fully demonstrated in Wang and Marongiu-Porcu (2015). The apparent matrix
permeability is the function of an interplay of three mechanisms: If geomechanical effects dominate, the matrix permeability will decrease with increasing depletion,
whereas if desorption and non-Darcy mechanisms dominate, the matrix permeability will increase with increasing depletion. The decoupled contribution on cumulative
production from the changes in hydraulic-fracture and natural-fracture properties caused by pressure/stress-dependent phenomena can be studied from Cases 3 and 4.

6
Cumulative Gas Production (MMcf)

3 Case 1

Case 2
2
Case 3
1
Case 4
0
0 500 1,000 1,500 2,000 2,500 3,000 3,500 4,000

Time (days)

Fig. 8—Comparison of cumulative gas production to show the decoupled effects of matrix, hydraulic fracture, and natural fracture (cases described in Table 4).

Case No. Description

Case 1 p wf = 500 psi; no pressure-dependent phenomena incorporated

Case 2 p wf = 500 psi; pressure-dependent phenomena: only matrix

Case 3 p wf = 500 psi; pressure-dependent phenomena: matrix and hydraulic fracture

Case 4 p wf = 500 psi; pressure-dependent phenomena: matrix, hydraulic fracture, and natural fracture

Table 4 — Descriptions of the different cases, results for which are shown in Fig. 8.

Effect of Flowing Bottomhole Pressure (BHP) on Long-Term Well Performance. We now consider several constant flowing-BHP conditions. The model setting
described in the preceding subsection is now run for two 10-year-production simulations under two drawdown scenarios, p wf ¼ 2,000 and 500 psi. Fig. 9 shows that the
more-aggressive drawdown strategy (i.e., lower p wf) performs better during the early life of the well, but eventually the production rate declines much faster compared
with the milder drawdown strategy (i.e., higher p wf). Consequently, the cumulative production curves cross (after approximately 3.5 years).
Cumulative Gas Production (MMcf)

67

345
p wf = 2,000 psi

012 p wf = 500 psi

0 500 1,000 1,500 2,000 2,500 3,000 3,500 4,000

Time (days)

Fig. 9—Cumulative production for two different drawdown scenarios.

This constitutes relevant evidence of the concept of pressure-drawdown management and validates the field-observed trends (as, for instance, in the soft
Haynesville Shale, where wells produced at very high initial rates tend to exhibit very rapid declines) that the penalty for lower initial production rates in unconventional
shale-gas reservoirs can yield substantially higher ultimate recovery.
In an attempt to find an optimal drawdown strategy that minimizes the pressure/stress-dependent effects and maximizes the cumulative recovery for the synthetic
case presented, we investigate then a set of drawdown strategies, spanning from p wf ¼ 2,000 to 500 psi.

10 2017 SPE Reservoir Evaluation & Engineering

ID: jaganm Time: 15:09 I Path: S:/REE#/Vol00000/170031/Comp/APPFile/SA-REE#170031


The results in Fig. 10 suggest that, excluding the obviously suboptimal scenario at p wf ¼ 500 psi and the too-conservative case at
p wf ¼ 2,000 psi, the optimal long-term drawdown strategy should be determined over imposing an economic optimization, as illustrated in Marongiu-Porcu et al. (2013).
Although not shown in Fig. 10, we run many iterative cases for various flowing BHPs, and conclude that the optimal p wf corresponding to maximum gas recovery for this
synthetic scenario is 1,750 psi. Nevertheless, it is important to emphasize the value of this integrated reservoir simulator, which enables execution of intensive parametric
studies aiming to determine whether the rapid production declines can be minimized with proper drawdown management to improve the ultimate cumulative recovery.

Cumulative Gas Production (MMcf)


4

p wf = 2,000 psi p wf = 1,500 psi


1
p wf = 1,000 psi p wf = 500 psi
0
0 500 1,000 1,500 2,000 2,500 3,000 3,500 4,000

Time (days)

Fig. 10—Cumulative production for the four different drawdown strategies considered.

Field-Data Example. In this final subsection, we present a production history match for the Marcellus Shale gas well previously studied in Eshkalak et al. (2014). Table 5 shows
all the input parameters used for the presented analysis; the additional parameters required as input in our model that do not appear in Table 5 are kept as the ones
previously reported in Table 3.

Input Parameters Value

Initial reservoir pressure p i 31 MPa

Wellbore pressure p wf 4.13 MPa

Intrinsic permeability measured at laboratory conditions k ∞0 50 nd

Hydraulic-fracture initial permeability k fi 2×10 3 md

Formation porosity measured at laboratory conditions φ 0 0.05

Fracturing spacing Y e 122 m

Reservoir thickness H 61 m

Drainage length parallel to the hydraulic fracture X e 152 m

Hydraulic-fracture half-length x f 152 m*

Total number of drainage units n 14

Langmuir volume constant V L 0.00624 m 3/ kg

Langmuir pressure constant P L 3.44 MPa

* Value not provided in Eshkalak et al. (2014)

Table 5 — Input parameters for the Marcellus Shale gas-well simulation. Data extracted from Eshkalak et al. (2014).

In Fig. 11, we compare the actual field data being matched by our production simulation that accounts for all the modeled pressure/ stress-dependent phenomena. In
addition, we present a simulation without accounting for pressure/stress-dependent phenomena (yellow curve), which clearly does not fit the field data. These two
simulations are better compared in Fig. 12, where we can observe a difference of approximately 20% in the cumulative production after the simulated forecast of 14
years, reflecting once again the deleterious effect of the depletion-related increase of the effective stress upon matrix- and fracture-flow potential.

Furthermore, it is relevant to notice from Fig. 11 that the early-time match of the model with the actual field data is relatively poor in the first 6 months. It is the
opinion of the authors that this mismatch may be primarily related to the flowback methodology used on the well. In fact, early-time production from shale reservoirs
often involves high drawdown in an effort to unload fracturing fluids and accelerate hydrocarbon production (i.e., from Table 5, the reported flowing BHP is 4.13 MPa
or 600 psi). In many cases, reduction of fracture conductivity during this stage can be attributed to rock failure at the fracture face and the resulting production of
formation material into the hydraulic fractures.

Deen et al. (2015) clearly described the numerous possible downhole-damage mechanisms that can affect connectivity to the reservoir and/or loss of fracture
conductivity, and how drawdown management during the early-time-flowback period is essential to minimize them. As more data confirming the negative effect of
aggressive drawdown become available, models and analysis methods are being developed to better understand and mitigate the risks associated with flowback and
better engineer this critical step of the life of the well.

2017 SPE Reservoir Evaluation & Engineering 11

ID: jaganm Time: 15:09 I Path: S:/REE#/Vol00000/170031/Comp/APPFile/SA-REE#170031


stress effects Model, with depletion-induced in-situ
stress-effects
Field data Model, without depletion-induced in-situ

Gas-Flow Rate (MMcf/D)


12

6 8 10

024

0 2 4 6 8 10 12 14 16

Time (years)

Fig. 11—Model validation with history match from field data (Eshkalak et al. 2014).

20

18
Cumulative Gas Production (MMcf)

16

14

12

10 Model, without depletioninduced in-situ


stress-effects
68
Model, with depletion-induced in-situ
24 stress effects

0
0 2 4 6 8 10 12 14 16

Time (years)

Fig. 12—Comparison of cumulative gas production for the two simulations presented in Fig. 11.

Reinicke et al. (2011) proposed the concept of mechanically induced fracture-face skin. According to their experiments and modeling work, the mechanical damage
at the rock/proppant interface begins immediately with loading the rock/proppant system and, for fracture-closure stresses less than 35 MPa, the damage is localized at
the fracture face. Microstructure analysis identified quartz-grain crushing, fines production, and pore-space blocking at the fracture face causing the observed
mechanically induced fracture-face skin.
To properly manage these critical issues, our integrated reservoir simulator could be fed with a flowing-BHP schedule generated by existing flowback simulators
(Reinicke et al. 2011; Suarez-Rivera et al. 2011) capable of providing the minimum flowing BHP corresponding to the critical (to prevent solids production from the
fracture face) drawdown as a function of production time and rate.
Our last comment concerns the comparison shown in Fig. 13 between the production history matches obtained with our model and the one presented in Eshkalak et
al. (2014), where an integrated reservoir model accounting for the pressure-dependent-property changes in the stimulated reservoir volume (rock matrix, induced
fractures, and hydraulic fractures) was used. Eshkalak et al. (2014) modeled the Klinkenberg (1941) and non-Darcy-flow effects in the matrix. However, they also
considered a simple exponential decline for the change in propped-hydraulic-fracture and induced-fracture permeability as a function of confining pressure. In turn, our
integrated-reservoirsimulator models explicitly include proppant crushing, proppant embedment, and creep deformation at the fracture face, and we showed that the
hydraulic-fracture permeability does not follow a simple exponential decline but is a convoluted function of rock properties, proppant characteristics, and proper proppant
placement within the hydraulically created fractures and the reactivated natural fractures.

data Our model


Eshkalak et al. (2014) model Field
Gas-Flow Rate (MMcf/D)

10 12

68

022

0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5

Time (years)

Fig. 13—Comparison of history matches performed with our model vs. the Eshkalak et al. (2014) model.

12 2017 SPE Reservoir Evaluation & Engineering

ID: jaganm Time: 15:09 I Path: S:/REE#/Vol00000/170031/Comp/APPFile/SA-REE#170031


Conclusion and Recommendations
The primary objective of this work is to address all the known causes of abnormal productivity decline in unconventional shale-gas formations and to develop a fully
coupled geomechanical/flow simulation model to predict and study these production conditions.
The model presented in this article is able to reproduce familiar field-observed trends, with impaired (and most likely noneconomical) long-term production
caused by higher and higher bottomhole drawdowns. Specifically, the effect of depletion-induced insitu-stress variations on short-term and long-term productivity is
quantified by taking into account several phenomena, such as stress-dependent matrix and natural-fracture permeability, as well as reduction in hydraulic-fracture
conductivity caused by proppant crushing, deformation, embedment, and fracture-face creep.

The fundamental conclusion of this work is that ignoring effects of any of the discussed phenomena results in overestimation of ultimate recovery. As such, our model
can be successfully used as a pressure-drawdown-management tool for recovery optimization in any given shale-gas formation. Proper drawdown management and the
penalty for lower initial production rates in unconventional shale-gas reservoirs can yield substantially higher ultimate recovery.

This integrated model can be used for optimization of key parameters during the hydraulic-fracture design, for fine tuning production history matching, and
especially as a predictive tool for pressure-drawdown management.
One element of improvement for the presented model will be the incorporation of early-time flowback-drawdown management. This would involve feeding our
integrated reservoir simulator with flowing-BHP schedules corresponding to the critical (to prevent solids production from the fracture face) drawdown as a function of
production time and rate.

Nomenclature
C/ ¼ Material constant
C f ¼ Fracture conductivity, L 3, m 3
CM ¼ Fracture conductivity multiplier
d f ¼ Compressibility of rock, Lt 2/ m, Pa 1
d m ¼ Diameter of absorbed gas molecules, L, m
DFN ¼ Discrete fracture network
E ¼ Young’s Modulus, m/Lt 2, Pa
F ¼ Net body force, m/Lt 2, Pa
G ¼ Shear Modulus, m/Lt 2, Pa
H ¼ Reservoir thickness, L, m
k 1 ¼ Intrinsic permeability, L 2, md
k 1 0 ¼ Intrinsic permeability at reference condition, L 2, md
k a ¼ Matrix apparent permeability, L 2, md
k nf,i ¼ Natural fracture permeability at initial reservoir condition, L 2, md
k nf ¼ Natural fracture permeability, L 2, md
k f ¼ Hydraulic fracture permeability, L 2, md
K n ¼ Knudsen number
M ¼ Average molecular weight, m/n, kg/mol
MPD ¼ Mean particle diameter, L, m
p c ¼ Net confining Pressure, m/Lt 2, Pa
p wf ¼ Flowing bottomhole pressure, m/Lt 2, Pa
P ¼ Reservoir Pressure, m/Lt 2, Pa
P c ¼ Critical pressure of mix gas, m/Lt 2, Pa
P i ¼ Initial reservoir pressure, m/Lt 2, Pa
P L ¼ Langmuir pressure, m/Lt 2, Pa
P 10 ¼ particle size corresponding to the P10 value on the cumulative proppant-size distributionrticle diameter, L, m
r ¼ Effective pore radius, L, m
r 0 ¼ Effective pore radius at reference condition, L, m
R ¼ Universal gas constant, mL 2/ t 2/ n/T, 8.3145 J/mol/K
R ed ¼ Roughness equivalent diameter, L, m
T ¼ Reservoir temperature, T, k
T c ¼ Critical temperature of mix gas, T, k
u i,j ¼ Component of displacement, L, m
V L ¼ Langmuir volume, L 3/ m, m 3/ kg
w ¼ Hydraulic fracture width, L, m
x f ¼ Hydraulic fracture length, L, m
X e ¼ Drainage length parallel to the hydraulic fractures, L, m
Y e ¼ Fractures spacing, L, m
a ¼ Biot’s coefficient
d i,j ¼ Kronecker delta
D PD ¼ Net reduction in fracture width due to proppant deformation, L, m
D PE ¼ Net reduction in fracture width due to proppant embedment, L, m
D w ¼ Total net reduction in fracture width, L, m
e( t) ¼ Time-dependent creep deformation, L, m
e ij ¼ Elastic strain
e kk ¼ Volumetric strain
k ¼ Characteristic length of flow path, L, m
¼ Poisson’s ratio
q m ¼ Matrix density, m/L 3, kg/m 3
r ij ¼ Effective stress, m/Lt 2, Pa
r m ¼ Mean effective stress, m/Lt 2, Pa
/ ¼ Matrix in situ porosity

2017 SPE Reservoir Evaluation & Engineering 13

ID: jaganm Time: 15:09 I Path: S:/REE#/Vol00000/170031/Comp/APPFile/SA-REE#170031


/ 0 ¼ Matrix porosity at reference condition
/ f ¼ Fracture porosity

References
Aguilera, R. 2002. Incorporating Capillary Pressure, Pore Throat Aperture Radii, Height Above Free-Water Table, and Winland r35 Values on Pickett
Plots. AAPG Bull. 86 ( 4): 605–624. https://doi.org/10.1306/61eedb5c-173e-11d7-8645000102c1865d.
Ajisafe, F., Thachaparambil, M., Lee, D. et al. 2016. Calibrated Complex Fracture Modeling Using Constructed Discrete Fracture Network from Seismic
Data in the Avalon Shale, New Mexico. Presented at the SPE Hydraulic Fracturing Technology Conference, The Woodlands, Texas, 9–11 February. SPE-179130-MS.
https://doi.org/10.2118/179130-MS.
API RP 61, Recommended Practices for Evaluating Short Term Proppant Pack Conductivity, first edition. 1989. Washington, DC: API. Barton, N., Bandish, S., and Bakhtar, K. 1985. Strength,
Deformation and Conductivity Coupling of Rock Joints. Int. J. Rock Mech. Min. 20 ( 3): 121–140. https://doi.org/10.1016/0148-9062(85)93227-9.

Barree, R. D., Cox, S. A., Barree, V. L. et al. 2003. Realistic Assessment of Proppant Pack Conductivity for Material Selection. Presented at the SPE An-
nual Technical Conference and Exhibition, Denver, 5–8 October. SPE-84306-MS. https://doi.org/10.2118/84306-MS. Berg, R. R. 1970. Method for Determining Permeability from Reservoir Rock
Properties. Trans. Gulf Coast Assoc. Geol. Soc. 20: 303–335. Bes ko¨k, A. and Karniadakis, G. 1999. Report: A Model for Flows in Channels, Pipes, and Ducts at Micro and Nano Scales. Microscale
Thermophys.
Eng. 3 ( 1): 43–77. https://doi.org/10.1080/108939599199864.
Blauch, M. E., Myers, R. R., Moore, T. R. et al. 2009. Marcellus Shale Post-Frac Flowback Waters —Where is All the Salt Coming From and What are
the Implications? Presented at the SPE Eastern Regional Meeting, Charleston, West Virginia, 23–25 September. SPE-125740-MS. https://doi.org/
10.2118/125740-MS.
Brannon, H. D., Malone, M. R., Rickards, A. R. et al. 2004. Maximizing Fracture Conductivity with Proppant Partial Monolayers: Theoretical Curiosity
or Highly Productive Reality? Presented at the SPE Annual Technical Conference and Exhibition, Houston, 26–29 September. SPE-90698-MS. https://doi.org/10.2118/90698-MS.

Britt, L. K., Smith, M. B., Klein, H. H. et al. 2016. Production Benefits from Complexity – Effects of Rock Fabric, Managed Drawdown, and Propped
Fracture Conductivity. Presented at the SPE Hydraulic Fracturing Technology Conference, The Woodlands, Texas, 9–11 February. SPE-179159-MS. https://doi.org/10.2118/179159-MS.

Brown, M. L., Ozkan, E., Raghavan, R. S. et al. 2011. Practical Solutions for Pressure-Transient Responses of Fractured Horizontal Wells in Unconven-
tional Shale Reservoirs. SPE Res Eval & Eng 14 ( 6): 663–676. SPE-125043-PA. https://doi.org/10.2118/125043-PA.
Chang, C. and Zoback, M. D. 2008. Creep in Unconsolidated Shale and its Implication on Rock Physical Properties. Presented at the 42nd US Rock
Mechanics Symposium (USRMS), San Francisco, 29 June–2 July. ARMA-08-130.
Chin, L. Y., Raghavan, R. S., and Thomas, L. K. 2000. Fully Coupled Geomechanics and Fluid-Flow Analysis of Wells with Stress-Dependent Perme-
ability. SPE J. 5 ( 1): 32–45. SPE-58968-PA. https://doi.org/10.2118/58968-PA.
Cho, Y., Ozkan, E., and Apaydin, O. G. 2013. Pressure-Dependent Natural-Fracture Permeability in Shale and its Effect on Shale-Gas Well Production.
SPE Res Eval & Eng 16 ( 2): 216–228. SPE-159801-PA. https://doi.org/10.2118/159801-PA.
Cipolla, C. L., Fitzpatrick, T., Williams, M. J. et al. 2011a. Seismic-to-Simulation for Unconventional Reservoir Development. Presented at the SPE Res-
ervoir Characterisation and Simulation Conference and Exhibition, Abu Dhabi, 9–11 October. SPE-146876-MS. https://doi.org/10.2118/146876-MS. Cipolla, C. L., Mack, M. G., Maxwell, S. C. et
al. 2011b. A Practical Guide to Interpreting Microseismic Measurements. Presented at the North Ameri-
can Unconventional Gas Conference and Exhibition, The Woodlands, Texas, 14–16 June. SPE-144067-MS. https://doi.org/10.2118/144067-MS. Cipolla, C. L., Weng, X., Mack, M. G. et al.
2011c. Integrating Microseismic Mapping and Complex Fracture Modeling to Characterize Hydraulic Frac-
ture Complexity. Presented at the SPE Hydraulic Fracturing Technology Conference, The Woodlands, Texas, 24–26 January. SPE-140185-MS. https://doi.org/10.2118/140185-MS.

Civan, F. 2010. Effective Correlation of Apparent Gas Permeability in Tight Porous Media. Transport Porous Med. 82 ( 2): 375–384. https://doi.org/
10.1007/s11242-009-9432-z.
Civan, F., Rai, C. S., and Sondergeld, C. H. 2011. Shale-Gas Permeability and Diffusivity Inferred by Improved Formulation of Relevant Retention and
Transport Mechanisms. Transport Porous Med. 86 ( 3): 925–944. https://doi.org/10.1007/s11242-010-9665-x.
Daneshy, A. A. 2004. Proppant Distribution and Flowback in Off-Balance Hydraulic Fractures. SPE Prod & Fac 20 ( 1): 41–47. SPE-89889-PA. https:// doi.org/10.2118/89889-PA.

Deen, T., Daal, J., and Tucker, J. 2015. Maximizing Well Deliverability in the Eagle Ford Shale Through Flowback Operations. Presented at the SPE An-
nual Technical Conference and Exhibition, Houston, 28–30 September. SPE-174831-MS. https://doi.org/10.2118/174831-MS. Dong, J.-J., Hsu, J.-Y., Wu, W.-J. et al. 2010. Stress-Dependence
of the Permeability and Porosity of Sandstone and Shale from TCDP Hole-A. Int. J.
Rock Mech. Min. 47 ( 7): 1141–1157. https://doi.org/10.1016/j.ijrmms.2010.06.019.
Eshkalak, M. O., Aybar, U., and Sepehrnoori, K. 2014. An Integrated Reservoir Model for Unconventional Resources, Coupling Pressure Dependent
Phenomena. Presented at the SPE Eastern Regional Meeting, Charleston, West Virginia, 21–23 October. SPE-171008-MS. https://doi.org/10.2118/ 171008-MS.

Far, M. E., Hardage, B., and Wagner, D. 2013. Inversion of Elastic Properties of Fractured Rocks from AVOAz Data Marcellus Shale Example. Proc.,
SEG Technical Program Expanded Abstracts 2013, 3133–3138. https://doi.org/10.1190/segam2013-0441.1.
Ghassemi, A. and Suarez-Rivera, R. 2012. Sustaining Fracture Area and Conductivity of Gas Shale Reservoirs for Enhancing Long-Term Production.
Final Research Partnership to Secure Energy for America Report 08100-48. Presented at the RPSEA Unconventional Gas Conference, Canonsburg, Pennsylvania, 17–18 April.

Goodman, R. E. 1974. The Mechanical Properties of Joints. Int. Soc. Rock Mech. 1: 127–140.
Guo, J. and Liu, Y. 2012. Modeling of Proppant Embedment: Elastic Deformation and Creep Deformation. Presented at the SPE International Production
and Operations Conference and Exhibition, Doha, Qatar, 14–16 May. SPE-157449-MS. https://doi.org/10.2118/157449-MS. Ilk, D., Jenkins, C. D., and Blasingame, T. A. 2011. Production
Analysis in Unconventional Reservoirs — Diagnostics, Challenges, and Methodologies.
Presented at the North American Unconventional Gas Conference and Exhibition, The Woodlands, Texas, 14–16 June. SPE-144376-MS. https:// doi.org/10.2118/144376-MS.

ISO 13503-5, Procedures for Measuring the Long-term Conductivity of Proppants. Petroleum and Natural Gas Industries – Completion Fluids and
Materials – Part 5, first edition. 2006. Geneva, Switzerland: International Organization for Standardization. Javadpour, F., Fisher, D., and Unsworth, M. 2007. Nanoscale Gas Flow in Shale Gas
Sediments. J Can Pet Technol 46 ( 10): 55–61. PETSOC-07-10-06. https://doi.org/10.2118/07-10-06.

Kaufman, P. B., Anderson, R. W., Parker, M. A. et al. 2007. Introducing New API/ISO Procedures for Proppant Testing. Presented at the SPE Annual
Technical Conference and Exhibition, Anaheim, California, 11–14 November. SPE-110697-MS. https://doi.org/10.2118/110697-MS.

14 2017 SPE Reservoir Evaluation & Engineering

ID: jaganm Time: 15:09 I Path: S:/REE#/Vol00000/170031/Comp/APPFile/SA-REE#170031


Klinkenberg, L. J. 1941. The Permeability of Porous Media to Liquid and Gases. API-41-200.
Lee, D. S., Elsworth, D., Yasuhara, H. et al. 2009. An Evaluation of the Effects of Fracture Diagenesis on Fracture Treatments: Modeled Response. Pre-
sented at the 43rd US Rock Mechanics Symposium (USRMS), Asheville, North Carolina, 28 June–1 July. ARMA-09-104. Lehman, L. V., Parker, M. A., Blauch, M. E. et al. 1999. Proppant
Conductivity —What Counts And Why. Presented at the SPE Mid-Continent Opera-
tions Symposium, Oklahoma City, Oklahoma, 28–31 March. SPE-52219-MS. https://doi.org/10.2118/52219-MS.
Liu, H.-H., Wei, M.-Y., and Rutqvist, J. 2013. Normal-Stress Dependence of Fracture Hydraulic Properties Including Two-Phase Flow Properties.
Hydrogeol. J. 21 ( 2): 371–382. https://doi.org/10.1007/s10040-012-0915-6.
Marongiu-Porcu, M., Economides, M. J., and Holditch, S. A. 2013. Economic and Physical Optimization of Hydraulic Fracturing. J. Nat. Gas Sci. Eng.
14 ( September): 91–107. https://doi.org/10.1016/j.jngse.2013.06.001.
McDaniel, G. A., Abbott, J., Mueller, F. A. et al. 2010. Changing the Shape of Fracturing: New Proppant Improves Fracture Conductivity. Presented at
the SPE Annual Technical Conference and Exhibition, Florence, Italy, 19–22 September. SPE-135360-MS. https://doi.org/10.2118/135360-MS. Merry, H., Ehlig-Economides, C. A., and Wei, P.
2015. Model for a Shale Gas Formation with Salt-Sealed Natural Fractures. Presented at the SPE An-
nual Technical Conference and Exhibition, Houston, 28–30 September. SPE-175061-MS. https://doi.org/10.2118/175061-MS. Miller, R. S., Conway, M., and Salter, G. 2010. Pressure-Dependent
Permeability in Shale Reservoirs Implications for Estimated Ultimate Recovery.
Oral presentation given at the AAPG Hedberg Conference, Austin, Texas, 5–10 December.
Offenberger, R., Ball, N., Kanneganti, K. et al. 2013. Integration of Natural and Hydraulic Fracture Network Modeling with Reservoir Simulation for an
Eagle Ford Well. Presented at the Unconventional Resources Technology Conference, Denver, 12–14 August. URTEC-1563066-MS. Okouma Mangha, V., Guillot, V. F., Sarfare, M. et al. 2011.
Estimated Ultimate Recovery (EUR) as a Function of Production Practices in the Haynes-
ville Shale. Presented at the SPE Annual Technical Conference and Exhibition, Denver, 30 October–2 November. SPE-147623-MS. https://doi.org/
10.2118/147623-MS.
Osholake, T., Wang, J. Y., and Ertekin, T. 2012. Factors Affecting Hydraulically Fractured Well Performance in the Marcellus Shale Gas Reservoirs.
J. Energy Resour. Technol. 135 ( 1): 013402. https://doi.org/10.1115/1.4007766.
Palmer, I. D., Moschovidis, Z. A., and Cameron, J. R. 2007. Modeling Shear Failure and Stimulation of the Barnett Shale after Hydraulic Fracturing. Presented
at the SPEHydraulic Fracturing Technology Conference, College Station, Texas, 29–31 January. SPE-106113-MS. https://doi.org/10.2118/106113-MS. Peyret, O., Drew, J., Mack, M. et al. 2012.
Subsurface to Surface Microseismic Monitoring for Hydraulic Fracturing. Presented at the SPE Annual Tech-
nical Conference and Exhibition, San Antonio, Texas, 8–10 October. SPE-159670-MS. https://doi.org/10.2118/159670-MS. Raghavan, R. and Chin, L. Y. 2002. Productivity Changes in
Reservoirs with Stress-Dependent Permeability. Presented at the SPE Annual Technical
Conference and Exhibition, San Antonio, Texas, 29 September–2 October. SPE-77535-MS. https://doi.org/10.2118/77535-MS. Reinicke, A., Bloecher, G., Zimmermann, G. et al. 2011.
Mechanically Induced Fracture Face Skin - Insights from Laboratory Testing and Numerical
Modelling. Presented at the SPE European Formation Damage Conference, Noordwijk, The Netherlands, 7–10 June. SPE-144173-MS. https:// doi.org/10.2118/144173-MS.

Sakhaee-Pour, A. and Bryant, S. 2012. Gas Permeability of Shale. SPE Res Eval & Eng 15 ( 4): 401–409. SPE-146944-PA. https://doi.org/10.2118/ 146944-PA.

Schubarth, S. and Milton-Tayler, D. 2004. Investigating How Proppant Packs Change Under Stress. Presented at the SPE Annual Technical Conference
and Exhibition, Houston, 26–29 September. SPE-90562-MS. https://doi.org/10.2118/90562-MS. Shi, X.-J., Li, C.-B., Adnan, A. et al. 2009. Viscoelastic Model with Variable Parameters for Earth
Media. Chinese J. Geophys. 52 ( 1): 33–39. https:// doi.org/10.1002/cjg2.1324.

Shoemaker, M., Zakhour, N., and Peacock, J. 2015. Integrating 3D Seismic and Geomechanical Properties with Microseismic Acquisition and Fracturing
Parameters to Optimize Completion Practices within the Wolfcamp Shale Play of the Midland Basin. Presented at the Unconventional Resources Technology Conference, San Antonio, Texas,
20–22 July. URTEC-2154184-MS. https://doi.org/10.15530/URTEC-2015-2154184. Sone, H. and Zoback, M. D. 2010. Strength, Creep and Frictional Properties of Gas Shale Reservoir Rocks.
Presented at the 44th US Rock Mechanics
Symposium and 5th US-Canada Rock Mechanics Symposium. Salt Lake City, Utah, 27–30 June. ARMA-10-463.
Suarez-Rivera, R., Deenadayalu, C., Chertov, M. et al. 2011. Improving Horizontal Completions on Heterogeneous Tight-Shales. Presented at the Cana-
dian Unconventional Resources Conference, Calgary, 15–17 November. SPE-146998-MS. https://doi.org/10.2118/146998-MS. Thachaparambil, M. V. 2015. Discrete 3D Fracture Network
Extraction and Characterization from 3D Seismic Data – A Case Study at Teapot Dome.
Interpretation 3 ( 3): 29–41. https://doi.org/10.1190/INT-2014-0219.1.
Valko´, P. P. and Lee, W. J. 2010. A Better Way to Forecast Production from Unconventional Gas Wells. Presented at the SPE Annual Technical Confer-
ence and Exhibition, Florence, Italy, 19–22 September. SPE-134231-MS. https://doi.org/10.2118/134231-MS.
Vega Navarro, O. G. 2012. Closure of Natural Fractures Caused by Increased Effective Stress, a Case Study: Reservoir Robore III, Bulo Bulo Field, Boli-
via. Presented at the SPE Latin America and Caribbean Petroleum Engineering Conference, Mexico City, 16–18 April. SPE-153609-MS. https:// doi.org/10.2118/153609-MS.

Wang, H. and Marongiu-Porcu, M. 2015. Impact of Shale-Gas Apparent Permeability on Production: Combined Effects of Non-Darcy Flow/Gas-Slip-
page, Desorption, and Geomechanics. SPE Res Eval & Eng 18 ( 4): 495–507. SPE-173196-PA. https://doi.org/10.2118/173196-PA.
Weng, X., Kresse, O., Chuprakov, D. et al. 2014. Applying Complex Fracture Model and Integrated Workflow in Unconventional Reservoirs. J. Pet. Sci. Eng. 124 ( December): 468–483.
https://doi.org/10.1016/j.petrol.2014.09.021.
Wilson, T. H., Hart, A. K., and Sullivan, P. 2014. Characterization of Marcellus Shale Fracture System for Fracture Model Development Using 3D Seis-
mic and Microseismic Data. Proc., SEG Technical Program Expanded Abstracts 2014, 2683–2687. https://doi.org/10.1190/segam2014-0547.1. Xiong, X., Devegowda, D., Villazon, G. G. M. et al.
2012. A Fully-Coupled Free and Adsorptive Phase Transport Model for Shale Gas Reservoirs
Including Non-Darcy Flow Effects. Presented at the SPE Annual Technical Conference and Exhibition, San Antonio, Texas, 8–10 October. SPE159758-MS. https://doi.org/10.2118/159758-MS.

Ypma, T. J. 1995 Historical Development of the Newton-Raphson Method. SIAM Rev. 37 ( 4): 531–551. https://doi.org/10.1137/1037125. Yu, W. and Sepehrnoori, K. 2014. An Efficient
Reservoir-Simulation Approach to Design and Optimize Unconventional Gas Production. J Can Pet
Technol 53 ( 2): 109–121. https://doi.org/10.2118/165343-PA.
Zakhour, N., Shoemaker, M., and Lee, D. 2015a. Integrated Workflow Using 3D Seismic and Geomechanical Properties with Microseismic and Stimula-
tion Data to Optimize Completion Methodologies: Wolfcamp Shale-Oil Play Case Study in the Midland Basin. Presented at the SPE Eastern Regional Meeting, Morgantown, West Virginia, 13–15
October. SPE-177298-MS. https://doi.org/10.2118/177298-MS. Zakhour, N., Sunwall, M., Benavidez, R. et al. 2015b. Real-Time Use of Microseismic Monitoring for Horizontal Completion
Optimization Across a
Major Fault in the Eagle Ford Formation. Presented at the SPE Hydraulic Fracturing Technology Conference, The Woodlands, Texas, 3–5 February. SPE-173353-MS.
https://doi.org/10.2118/173353-MS.
Zhang, Z. and Li, X. 2016. The Shear Mechanisms of Natural Fractures during the Hydraulic Stimulation of Shale Gas Reservoirs. Materials 9 ( 9): 713–727. https://doi.org/10.3390/ma9090713.

2017 SPE Reservoir Evaluation & Engineering 15

ID: jaganm Time: 15:09 I Path: S:/REE#/Vol00000/170031/Comp/APPFile/SA-REE#170031


Ziarani, A. S. and Aguilera, R. 2012. Knudsen’s Permeability Correction for Tight Porous Media. Transport Porous Med. 91 ( 1): 239–260. https:// doi.org/10.1007/s11242-011-9842-6. Zienkiewicz, O.
C. and Taylor, R. L. 2005. The Finite Element Method: Vol. 1: The Basis, fifth edition. London: Elsevier.

Ankit Mirani recently joined Essar Oil Limited in the role of production engineer in West Bengal, India. His main interests are production engineering, hydraulic fracturing, production optimization, and
unconventional-reservoir development. Mirani holds a master’s degree in petroleum engineering from the University of Houston.

Matteo Marongiu-Porcu is a senior completions consultant with Schlumberger. He has 15 years of experience in upstream petroleum-production engineering and operations, hydraulic fracturing,
matrix acidizing, completions, conventional and unconventional production-enhancement technologies, reservoir engineering, pressure and production transient analysis, and engineering
management. Furthermore, for the past 4 years, Marongiu-Porcu has held concurrent adjunct-professor appointments at the University of Houston. Previously, he worked for the Italian national oil
company Eni E&P as a production engineer and has been an associate partner and senior consultant for Economides Consultants. Marongiu-Porcu holds a 5-year degree in chemical engineering
from Politecnico di Milano, Italy; a master’s degree in petroleum engineering from the University of Houston; and a PhD degree in petroleum engineering from Texas A&M University.

HanYi Wang is a PhD degree candidate in the Department of Petroleum and Geosystems Engineering at the University of Texas at Austin. Previously, he worked for Schlumberger. Wang’s research
is mainly focused on reservoir modeling, hydraulic fracturing, well performance, production optimization, and unconventional-reservoir development. He has years of industry research and consulting
experience. Wang serves as technical editor/reviewer for several international journals. He holds a master’s degree in petroleum engineering from the University of Houston and a bachelor’s degree
with honors in gas storage and transportation engineering from Southwest PetroleumUniversity, China.

Philippe Enkababian is currently the Schlumberger worldwide stimulation domain manager, responsible for maximizing the value of stimulation technology and engineering for the industry. He has
worked with Schlumberger since 1994, where he has held numerous positions including technical support engineer, production and stimulation engineer, and stimulation domain manager. More
recently, Enkababian was the Russia and central Asia marketing and technology manager for the Schlumberger Production Group. He has 23 years of experience in upstream petroleum-production
engineering and operations, hydraulic fracturing, matrix acidizing, water and scale control, conventional and unconventional completions and production-enhancement technologies, and engineering
management. Enkababian holds a bachelor’s degree in mechanical engineering from the E´cole Nationale Supe´ rieure d’Arts et Me´ tiers de Paris.

16 2017 SPE Reservoir Evaluation & Engineering

View publication
publication stats stats View
ID: jaganm Time: 15:09 I Path: S:/REE#/Vol00000/170031/Comp/APPFile/SA-REE#170031

Anda mungkin juga menyukai